<<

THE REGULATION OF SECRETORY CLUSTERIN EXPRESSION

AFTER IONIZING RADIATION EXPOSURE

By

TRACY CRISWELL

Submitted in partial fulfillment of the requirements

For the degree of Doctor of Philosophy

Thesis Advisor: Dr. David A. Boothman

Program in Molecular and Cellular Basis of

Department of Pathology

CASE WESTERN RESERVE UNIVERSITY

May 2004 CASE WESTERN RESERVE UNIVERSITY

SCHOOL OF GRADUATE STUDIES

We hereby approve the dissertation of

______

candidate for the Ph.D. degree *.

(signed)______(chair of the committee)

______

______

______

______

______

(date) ______

*We also certify that written approval has been obtained for any proprietary material contained therein. Dedication

This thesis is dedicated to my family, especially my husband, Jamie, my children,

McKenzie and Gaelen and my parents, Linda and Thomas Flamm. I would not be the

person I am today without their perpetual love and support. 2 Table of Contents

Title Page 1

Signature Sheet 2

Dedication 3

Table of Contents 4

List of Tables 5

List of Figures 6-7

Acknowledgements 8

List of Abbreviations 9-10

Abstract 11-12

CHAPTER 1: Introduction

1.1: Clusterin 13-21

1.2: Transcription Factors Activated in Mammalian Cells

After Clinically Relevant Doses of Ionizing Radiation 22-56

CHAPTER 2: Repression of IR-Inducible Clusterin

Expression by the Tumor Suppressor 57-88

CHAPTER 3: Induction of Clusterin, a Pro-Survival Factor,

Requires a Delayed Activation of the IGF-1R/MAPK

Signal Cascade After IR 89-124

CHAPTER 4: Conclusions and Future Directions 125-134

BIBLIOGRAPHY 135-186

3 List of Tables

Table 1.1. Various names assigned to clusterin

Table 1.2. IR-inducible transcription factors

Table 2.1. sCLU is a stress protein induced by a variety of cytotoxic agents

Table 2.2. Effect of p53 status on sCLU basal and IR-inducible expression

4 List of Figures

Figure 1.1. Schematic of clusterin structure

Figure 1.2. A schematic diagramming the processing of the CLU message into the

various CLU protein forms

Figure 1.3. DNA damage resulting from ionizing radiation exposure results in activation

of signaling cascades that lead to either survival or

Figure 1.4. IR stimulates the phosphorylation of IKK, which results in the degradation of

IκB allowing for NF-κB to move into the nucleus and activate downstream

Figure 2.1. sCLU is transcriptionally upregulated after IR exposure in MCF-7 human

breast cells

Figure 2.2. sCLU basal levels are elevated in MCF-7 and RKO cells that over-express the

HPV E6 protein

Figure 2.3. sCLU is induced in HCT116:p53-/- cells, but not in p53-/- HCT116 parental

cells

Figure 2.4. sCLU is not induced in HCT116:p21-/- cells

Figure 2.5. sCLU is not cell cycle regulated

Figure 3.1. siRNA to sCLU in MCF-7 cells results in radio-sensitization

Figure 3.2. Activation of MAPK by IR

Figure 3.3. EGFR is not an upstream activator of sCLU

5 Figure 3.4. Inhibition of Insulin-like Growth Factor-1 Receptor (IGF-1R) abrogates

sCLU induction after IR

Figure 3.5. c-Src is an upstream activator of sCLU induction after IR

Figure 3.6. Activation of the MAPK cascade is required for sCLU induction after IR

Figure 3.7. The Egr-1 transcription factor can bind to the CLU promoter and is required

for promoter activation after IR

Figure 4.1. Model indicating hypothetical cross-talk between the signals for sCLU

induction after IR and p53 repression.

6 Acknowledgements

I would like to acknowledge everyone who has helped me through the past several years during my graduate career.

First I would like to thank my mentor, David A. Boothman, Ph.D who has guided me when I needed it and left me to find my own path when I needed to do that as well.

You encouraged me from the start and introduced me to a scientific community that I would not have found on my own. I would also like to thank Dr. Lindsey Mayo who has guided me through the clusterin signaling maze over the past year.

I would also like to thank my committee members, Dr. Alan Tartakoff, Dr. Sanjay

Pimplikar, Dr. Susann Brady-Kalnay, Dr. Lindsey Mayo and Dr. Clive Hamlin for helping me to “stay on the straight and narrow” path. Their guidance helped to keep me focused and their support helped me to become the scientist that I am today.

Thank you to everyone in the Boothman lab, both past and present. You have been my company and my friends for the past several years, and the relationships I have made have helped me through the rough times.

Lastly, I want to thank my husband, Jamie, and my children, McKenzie and

Gaelen. Thank you for your never ending support and love. Thank you for putting up with my work on the weekends and evenings. Thank you for lifting me up when I was low and letting me rest when I was tired, and thank you for being there every step of the way.

7 List of Abbreviations

ATM ataxia telengiectasia mutated

ATR ataxia telengiectasia related gene

BASC BRCA-containing assembly complex bp

CDK cyclin-dependent kinase

CLU clusterin

CMV cytomegalovirus dn dominant negative

DSB DNA double-strand break

EGF epidermal growth factor

EGFR epidermal growth factor receptor

EMSA electrophoretic mobility shift assay

ER endoplasmic reticulum

FBS fetal bovine serum

HUVEC human umbilical vein endothelial cells

IEGs immediate early genes

IGF-1 insulin-like growth factor-1

IGF-1R insulin like growth factor-1 receptor

IGFBPs insulin-like growth factor binding

IR ionizing radiation

MAPK mitogen activated protein kinase

8 NAC n-acetylcysteine nCLU nuclear clusterin

NLS nuclear localization sequence

PDT photodynamic therapy

PI propidium iodide psCLU pre-secretory clusterin

RCEs retinoblastoma control elements

RCPs retinoblastoma control proteins

ROS reactive oxygen species

SAPK stress activated protein kinase sCLU secretory clusterin

SERCA sarco-endoplasmic reticulum calcium ATPase

Src CA constitutively active Src

Src KD kinase dead Src

SREs serum response elements

SSBs DNA single-strand breaks

TG thapsigargin

TPA tetradecanoylphorbol acetate

UV ultraviolet radiation

VWF von Willebrand factor wt wild-type xip x-ray induced protein

9 The Regulation of Secretory Clusterin Expression after Ionizing

Radiation Exposure

Abstract

By

TRACY CRISWELL

Radiation therapy is a common treatment for many types of tumors. Therefore, it is vital to understand the cellular responses to radiotherapy in malignant cells, as well as the surrounding normal tissues for optimizing the efficacy of this therapy. Our laboratory identified clusterin (CLU) as a protein/transcript that could be induced by doses of ionizing radiation (IR) as low as 0.02 Gy, suggesting a role for CLU in the cellular response to IR. CLU is a secreted that has been implicated in many normal biological processes as well as many pathological processes (Chapter 1). While the exact function of CLU still remains unknown, it has been suggested that secretory CLU

(sCLU), the fully processed and glycosylated form of CLU, plays a role in cytoprotection after cellular stress.

This thesis characterizes the regulation of sCLU after IR exposure. We observed that sCLU protein levels were low and only marginally induced after IR in cell lines containing functional wild-type p53. In contrast, cell lines expressing mutant p53 or null p53 had high basal levels of sCLU. This led us to investigate the role of p53 as a transcriptional repressor of sCLU (Chapter 2). Using genetically matched cell lines, we demonstrated that sCLU is transcriptionally repressed by p53.

10 Further work was done to characterize the signaling pathways required for sCLU induction after IR. We demonstrated that sCLU is involved in cell survival after IR using sCLU specific siRNA (Chapter 3). Furthermore, we demonstrated that the Src-Mek-Erk signaling cascade is reactivated 24-72 h after IR, which corresponds to the late induction of sCLU, and that the signal for sCLU induction after IR is generated by activation of the insulin-like growth factor-1 receptor (IGF-1R). Finally, we demonstrate that this signaling cascade culminates in the activation of Egr-1, which is required for the induction of sCLU after IR.

11 Chapter 1: Introduction

1.1 Clusterin

Clusterin (CLU) is a secreted glycoprotein originally identified in 1983 as a protein involved in cellular aggregation in ram rete testis fluid (1, 2). In 1987, Leger et al. demonstrated that CLU was induced in the ventral rat prostate after castration and named it testosterone-repressed prostate message 2 (TRPM-2) (3). Since then, many groups working in diverse fields have identified/cloned CLU, resulting in diverse nomenclature for this protein (Table 1.1) (1, 3-11). CLU has been implicated in many physiological processes such as metabolism (12, 13), complement regulation (10,

14, 15), cell differentiation (16, 17), reproduction (1, 18, 19), tissue remodeling/regeneration (20-25), (26-28) and cell death (29-35). CLU has also been implicated in many diverse disease processes including (36-38),

Alzheimer’s disease (39-41), (24, 42, 43), preeclampsia (44), lupus erythematosus (45), retinitis pigmentosa (46, 47) and scrapie (29).

Elevated levels of secretory CLU (sCLU) protein and mRNA have been observed in several different types of human neoplasias and malignancies including

(48-50), (51), hemangiomas (52), malignant lymphomas (53-55), ovarian cancer (56, 57), (58), colorectal carcinoma (59) and renal clear cell carcinomas (60, 61). Additionally, it has been shown that forced over-expression of sCLU in transformed cell lines resulted in an increased resistance to doxorubicin, cisplatin and taxol (48, 62, 63) and abrogation of CLU mRNA expression following

12 Table 1.1

Various names assigned to clusterin and the species from which it was isolated.

Name Species Sample Type Association Year Ref Clusterin ram rete testes fluid reproduction 1983 (1, 2) Glycoprotein III (GP-III) bovine adrenal medulla chromaffin granules 1984 (6) Sulfated glycoprotein-2 (SGP-2) rat sertoli cells reproduction 1984 (64) Testosterone-repressed prostate message 2 (TRPM2) rat prostate apoptosis 1987 (3) gp80 canine renal cells vectorial 1987 (7) SP-40,40 human serum complement regulation 1988 (18) pTB16 human brain gliomas and epileptic foci 1988 (65) T64 quail neuroretinal cells cell transformation 1989 (66) complement lysis inhibitor (CLI) human serum complement regulation 1989 (9) SGP-2 hamster brain scrapie 1989 (20) pADHC-9 human brain Alzheimer’s disease 1989 (67) X-ray induced protein-8 (XIP-8) human cells IR induction 1989 (11) J (Apo-J) human blood lipid transport 1990 (68) Clusterin rat spermatozoa reproduction 1991 (8, 18) NA1/NA2 human blood lipid transport 1991 (8) K611 human retina Retinitis pigmentosa 1992 (69) pc38k porcine vascular smooth muscle nodule formation 1992 (70) binding protein 1 (KUB 1) human breast epithelial cells Ku70 binding protein 1999 (71)

13 antisense expression lead to modest chemo- and IR-sensitization in various cell lines (62,

72-74). These data support a cytoprotective role for sCLU and suggest that over- expression of endogenous CLU provides a survival advantage for the tumors in which it is expressed.

The function of CLU still remains to be elucidated. The defining property of

CLU seems to be that it is induced under conditions that result in cellular stress. This fits with its proposed role as a molecular involved in the clearance of cell debris after cellular stress (75-78). Additionally, Humphreys et al., demonstrated that CLU has properties similar to small heat shock proteins (75), and that sCLU can protect the proteins S-transferase and catalase from heat-induced precipitation, due to improper , through the formation of high molecular weight complexes (75).

Furthermore, Poon et al. demonstrated that the ability of CLU to protect cells from precipitation was maximal at a slightly acidic pH (79). This suggests a role for sCLU at sites of injury and , where the pH is slightly acidic to prevent against infection.

The development of a CLU knock-out mouse has allowed further insight into the function of this protein (80). The CLU knock-out mice develop a more severe myocarditis, compared to wild-type mice, with substantial scarring after challenge with murine myosin (80), and Han et al., demonstrated that CLU deficient mice had 50% less cell death after hypoxic/ischemic insult (81). Additionally, Rosenberg et al., showed that

CLU deficient mice had increased immune complex deposition in the kidneys, indicative of progressive glomerulopathy (82). These data suggest a protective role for CLU during tissue injury and inflammatory responses. Additionally, CLU (also termed

14 apolipoprotein J) and apolipoprotein E (ApoE) double knock-out mice showed a decrease in deposition of fibrillar β-amyloid in the brain (83), supporting a role for CLU in

Alzheimer’s disease and of aging.

The signaling pathways that result in CLU induction after stress have not been elucidated. In 1992, Herrault et al., found that v-Src could induce transcription of the avian CLU gene (69). The involvement of Src in human CLU induction has not been investigated. TGF-β also appears to play a role in CLU gene and protein expression (84,

85), potentially through c-Fos binding to an Ap-1 site located in the CLU promoter (86).

Additionally, extracellular CLU can bind to the TGF-β receptor II, suggesting a possible feedback mechanism where CLU affects the downstream TGF-β signaling pathways.

The role of CLU in TGF-β mediated signaling pathways has not been fully elucidated.

B-Myb has also been shown to bind and activate the CLU promoter (87), but the physiologic relevance of this has yet to be determined.

NF-κB has also been shown to regulate sCLU expression. Saura et al., demonstrate that NF-κB is required for LPS stimulation of sCLU expression in glial cells

(88) and Li et al., using microarray technology, demonstrate that TNF-stimulated CLU expression is dependent on the NF-κB/IKK complex (89). Of interest, Santilli et al., demonstrate that sCLU can disrupt NF-κB signaling by stabilizing the inhibitors of NF-

κB (IκBs) (90). This suggests a possible negative feedback loop, where NF-κB stimulates the expression of sCLU after LPS or TNF-α exposure that then acts to stabilize IκBs and silence NF-κB signaling. Signaling after IR has not been elucidated.

Our laboratory isolated CLU as an X-ray-inducible protein/transcript (xip8), in which northern blot and nuclear run-on studies demonstrated enhanced transcript

15 synthesis and steady state mRNA accumulation within IR-exposed human malignant melanoma cells (11, 91, 92). We further identified CLU as a Ku70 binding partner (KUB

1) through a yeast two hybrid screen using Ku70 as bait. Since Ku70 is predominantly localized to the nucleus, we proposed a nuclear form of the CLU protein (nCLU). We found that a 49 kDa nCLU protein resided in the of unirradiated MCF-7 cells and that after IR, this form translocated to the nucleus as a 55 kDa protein (35). The post-translational modifications that result in this change in molecular weight are currently being investigated.

The human CLU mRNA contains two AUG start sites separated by 32 amino acids – therefore one message appears to encode two separate proteins, whose processing is dictated by the presence (sCLU) or absence (nCLU) of this leader peptide sequence

(19, 93) (Figure 1.1). When the CLU mRNA is read at its first AUG sequence, a leader peptide targets the protein to the endoplasmic reticulum (ER). This 60 kDa unmodified peptide (pre-sCLU) is cleaved at an α/β cleavage site to yield two 40 kDa subunits that heterodimerize through the formation of five disulfide bonds to form the mature 80 kDa secretory form of the protein that is further processed and glycosylated in the (Figure 1.2) (5, 19). sCLU separates as an 80 kDa protein or ~40 kDa proteins under SDS-PAGE non-reducing or reducing conditions, respectively (35). Treatment of log-phase MCF-7 cells with ≥ 2 cGy (2 rads) of ionizing radiation (IR) results in dramatic increases in production of sCLU message and protein, as well as a massive accumulation of the 60 kDa peptide that is presumably present in the ER and golgi of IR-exposed cells within 24-72 h post-IR exposure (35, 94).

16 ATG ATG

NLS N C Coiled-coil Coiled-coil domain domain aa 42-98 aa 319-349 α/β cleavage site

Figure 1.1. Schematic of CLU gene structure. Red hashed boxes correspond to the two coiled –coil domains at amino acids 42-98 and 319-349 respectively (95). Black boxes correspond to two potential nuclear localization sequences.

17 We recently noted that, due to of the CLU mRNA, CLU can be translated from the second AUG start site, resulting in the formation of a 49 kDa pre- nuclear (pre-nCLU) protein observed in the cytoplasm of control non-irradiated MCF-7 human breast cancer cells (95). The nCLU protein contains two putative nuclear localization signals (NLS); the first one located after the second AUG start site and the second located at the C-terminus (Figure 1.1). We propose that these NLSs are kept concealed either through homodimerization or protein folding, through the two coiled- coil domains, until after cellular stress (95). After treatment of log-phase MCF-7 cells with as little as 1.0 Gy of IR, the levels of pre-nCLU protein, as well as a ~55 kDa nuclear form of the protein (nCLU), dramatically increased 48 to 72 h (35, 71). Analyses of nCLU (isolated from the nuclei of IR-exposed MCF-7 cells) revealed that the protein was not cleaved at its α/ß site, since its migration was not altered under reducing or non- reducing SDS-PAGE conditions. Furthermore, accumulation of nCLU protein appears to be sufficient for signaling cell death, even in the absence of IR exposure, and the protein appears to associate with the Ku70/Ku80 DNA double strand break repair machinery

(35).

18 NLS

Clusterin AUG AUG mRNA

Leader Peptide NLS 60 kDa Unmodified α β Precursor α/β cleavage Form (psCLU)

80 kDa

α S S S S S Modified S S S S S Secretory β Form (sCLU)

Figure 1.2. The human CLU mRNA contains two AUG start sites separated by 32 amino acids. When the CLU mRNA is translated from its first AUG sequence, a leader peptide targets the protein to the endoplasmic reticulum (ER) as it is being translated. This 60 kDa unmodified peptide (psCLU) is cleaved at an α/β cleavage site to yield two 40 kDa subunits that heterodimerize through the formation of 5 disulfide bonds to form the 80 kDa secretory form of the protein that is highly glycosylated in the golgi apparatus. Translation from the second AUG start site results in the production of the nuclear form of CLU (nCLU).

19 In summary, CLU is a protein implicated in many normal biological processes and many pathological disease processes. The function of sCLU still remains to be elucidated, but it seems to play a role in cellular stress responses. sCLU appears to provide cytoprotection against cellular injury and inflammatory responses while the nuclear form of the protein appears to be cytotoxic. The regulation of this protein after stress is not understood. A better understanding of this protein and its various roles in cellular responses to stress will allow us to generate better treatments and therapies for many of these different pathological processes.

20 1.2: Transcription factors activated in mammalian cells after clinically relevant

doses of ionizing radiation

This work was published in Oncogene

Oncogene 2003 Sept 1; 22; 5813-5827

Abstract

Over the past fifteen years, a wealth of information has been published on transcripts and proteins ‘induced’ (requiring new synthesis) in mammalian cells after ionizing radiation

(IR) exposure. Many of these studies have also attempted to elucidate the transcription factors that are ‘activated’ (i.e., not requiring de novo synthesis) in specific cells by IR.

Unfortunately, all too often this information has been obtained using supra-lethal doses of

IR, with investigators assuming that induction of these proteins, or activation of corresponding transcription factors, can be “extrapolated” to low-dose IR exposures.

This chapter focuses on what is known at the molecular level about transcription factors induced at clinically relevant (<2 Gy) doses of IR. A review of the literature demonstrates that extrapolation from high to low doses of IR is inaccurate for most transcription factors and most IR-inducible transcripts/proteins, and that induction of transactivating proteins at low doses must be empirically derived. The pathways stimulated after high versus low doses of IR, which act to transactivate certain transcription factors in the cell, will be discussed. To date, only three transcription factors appear to be responsive (i.e., activated) after physiological doses (doses wherein cells survive or recover) of IR. These are p53, NF-κB and the Sp1- related retinoblastoma control proteins (RCPs) and related Egr-1 transcription factors.

21 Clearly, more information on transcription factors and proteins induced in mammalian cells at clinically or environmentally relevant doses of IR is needed to understand the role of these stress responses in cancer susceptibility/resistance and radio- sensitivity/resistance mechanisms.

IR-inducible responses in mammalian cells

Ionizing radiation exposure of mammalian cells causes a spectrum of lesions within the cell. At the DNA level, these lesions include DNA single- and double-strand breaks (SSBs, DSBs, respectively), DNA base damage and apyrimidinic/apurinic sites and DNA-protein cross-links. Although the most important lesion for triggering cell death is the formation of DSBs, other DNA lesions (e.g., formation of 8-hydroxyguanine) may have dramatic mutagenic and carcinogenic, consequences (96). Aside from DNA damage, IR also causes a spectrum of other lesions in cellular macromolecules (e.g., lipid peroxidation) due to the generation of reactive oxygen species (ROSs) created by the ionization of water and iron-related Fenton reactions within the cell. Although these non-

DNA lesions are probably not lethal to the cell, they can stimulate various signal transductions pathways (such as protein kinase C (PKC), JNK, ceramide, and MAPK activation) after certain doses of IR. Thus, IR causes a spectrum of DNA and non-DNA lesions that represent potential signals that activate sensory proteins (97). These sensory proteins act to signal that damage has occurred in the cell, as well as regulate processes that both halt cell cycle progression and stimulate repair of DNA lesions.

22 Signals from DNA

At the DNA level, damage sensors stimulated by IR include: (a) the DNA- dependent protein kinase repair complex (DNA-PK, composed of Ku70, Ku80 and DNA-

PK catalytic subunit [DNA-PKcs]) (98, 99); (b) the ataxia telengiectasia (AT) mutated gene (ATM)/c-ABL complex (100, 101); (c) the ataxia telangiectasia-related gene (ATR)

(102, 103); (d) the NBS1-MRE-11-Rad50 complex (104); (e) the DNA mismatch repair complex (97, 105-107); and (f) the Rad9-Hus1-Rad-1 (9:1:1) complex (108-110), among others as yet to be defined processes within mammalian cells. It may be possible that many of these damage sensors exist in a huge complex, referred to as the BRCA- containing assembly complex (BASC complex), and are activated based on the type and/or extent of DNA damage (111) (Figure 1.3).

These damage sensors not only play a role in DNA repair, but also simultaneously regulate cell cycle checkpoint responses that control life and death decisions. Eventually, the damage sensors stimulate specific signal transduction processes that activate specific transcription factors, such as Sp-1, p53, and NF-κB. These activated transcription factors, in turn, control gene expression, cell cycle checkpoint regulation, and under severe conditions may stimulate apoptosis. In recent years, our knowledge of the circuitry of these pathways has increased dramatically, mostly from the development of sound genetic models. We are only now beginning to understand the significance of these pathways in cell cycle progression, apoptosis, survival, mutagenesis and carcinogenesis (particularly when these pathways are altered due to accumulated or inherited mutations). However, the promise of completely understanding the retrograde signaling processes that emerge from DNA damage to cell cycle checkpoint control is

23 tempered by the fact that much of our information has been obtained to date, using rather high (most times supra-lethal) doses of agents, such as IR. In this chapter, we will examine the transcription factors activated after IR, with particular emphasis on doses required for their activation. For our discussion, ‘activation’ refers to conversion of an inert protein to an active form that can stimulate transcription of a gene. Expression of the new transcript or protein as a result of the ‘activation’ of these transactivation factors is an ‘induced’ gene product, such as the growth arrest and damage inducible (GADD) gene products (112) or x-ray–induced transcripts/proteins (XIPs) (91). We will focus on the activation processes of transcription factors following low, clinically relevant doses of

IR.

Signals from non-DNA targets

Much less is understood about IR-induced damage signaling from non-DNA targets that eventually cause transactivation of specific proteins (i.e., transcription factors) at clinically relevant doses of IR. Most of the studies attempting to elucidate these signaling events from non-DNA targets have utilized supra-lethal doses of IR and the triggering mechanisms have, in general, not been elucidated. Specifically, exposure of cells to IR results in a spectrum of lesions in cell membranes (e.g., lipid peroxidation).

Lipid peroxidation can result in the activation of PKC, MAPK, casein kinase, the ceramide pathway, JNK, ERK and other signal transduction pathways. In this way, damage from cellular plasma membranes is signaled inward to activate specific transcription factors. As stated above, these pathways have not been, in general, evaluated at low doses of IR, and the relevance of such signal transductions in the

24 Figure 1.3. DNA damage resulting from ionizing radiation exposure results in activation of signaling cascades that lead to either cell survival or apoptosis.

25 activation of specific transcription factors will be evaluated below for each known IR- stimulated transcription factor. Since IR causes simultaneous damage to DNA and non-

DNA targets, it can be a daunting task to evaluate the relative strength of DNA versus non-DNA targets for triggering a given signal transduction process. This, in turn, makes it difficult to determine the consequences of a particular activated signal transduction and/or transcription factor pathway in the overall response of the cell to IR. In particular, the use of high doses of IR to stimulate signal transduction, transcription factor activation, cell cycle checkpoint, or apoptotic responses in a more rapid timeframe (a rationale used by many investigators for the use of supra-lethal IR doses) appears to create many simultaneous (and at times conflicting or synergistic) signals that are clearly not applicable to clinically relevant lower doses of IR. These issues will be addressed in our discussion of known IR-activated transcription factors below.

Known IR-activated transcription factors

A review of the literature indicates a fairly select number of transcription factors activated by IR at any dose tested and in a variety of cell models. This list includes p53, nuclear factor-kappa-B (NF-κB), the retinoblastoma control proteins (RCPs), the early growth response 1 transcription factor (Egr-1, an Sp-1-like factor), the c-Fos/c-Jun AP-1 complex, and the octamer-binding protein-1 (Oct-1) transcription factor (Table 1.2). The transcription factors in Table 1 represent the known transcription factors activated by IR as found by a review of the current literature. In the sections to follow, we will discuss the data that demonstrates the IR activation of these transcription factors. We will then explore whether these transcription factors are activated at low clinically relevant doses

26 of IR and discuss their proposed functional significance in IR responses in normal or tumor cells.

IR stimulation of p53 transcriptional responses

The p53 tumor suppressor gene is mutated in ~50% of all human tumors and has been extensively reviewed (113-116). This review will specifically focus on the activation and functions of p53 by IR. Most of the studies investigating these functions have been done at high doses of IR (>4 Gy), while the role of p53 in low-dose responses still remains to be clarified. The p53 protein is tightly regulated and remains at low levels in unstressed cells (117), but is rapidly stabilized by various types of cellular stresses, including ionizing (IR) and ultraviolet (UV) radiations, hypoxia, stabilization/destabilization of microtubules, and ribonucleotide depletion (115, 118-120).

In response to DNA damage, p53 is rapidly stabilized through post-translational modifications that may include phosphorylation, acetylation, sumoylation, glycosylation and ribosylation. In general, IR is believed to activate (stabilize) p53 mainly through phosphorylation by the ATM kinase (113, 121, 122).

p53 plays a critical role in maintaining genomic integrity after cellular stress by acting as a transcription factor for a number of downstream genes that may mediate either cell cycle arrest or apoptosis, thereby preventing propagation of damaged DNA. The response to p53 stabilization is cell-type dependent, as well as damage dependent. For example, epithelial cells tend to undergo growth arrest after IR exposure, whereas lymphoid cells tend to undergo p53-dependent apoptosis (123). The signals that

27 Table 1.2

IR-inducible transcription factors.

Induction Transcription IR Dose Cell Type Used Time Reference Factor Range (Gy) Course (h) lymphocytic, p53 epithelial cancer 0.2-20 2-4 (124, 125) cells lymphocytic, NF-κB epithelial cancer 0.5-20 0.5-2 (126) (p50/p65) cells melanoma, Sp1 (RCPs) squamous 0.1-10 2-10 (127, 128) cell carcinoma lymphocytic, Egr-1 epithelial cancer 4-20 0.3-4 (129) cells lymphocytic, AP-1 epithelial cancer 2-20 1-5 (130, 131) cells, keratinocytes epithelial cancer Oct-1/NF-Y 5-20 4-8 (132) cells, keratinocytes

28 determine whether a cell undergoes cell cycle arrest or p53-dependent apoptosis after

DNA damage still remain to be elucidated.

The p53 protein consists of four major domains, a transactivation domain at the

N-terminus, a proline-rich domain, a DNA-binding domain and a carboxy-terminal tetramerization domain (133, 134). The transactivation domain contains the binding site (135), as well as binding sites for members of the transcription machinery.

The proline-rich domain is important for efficient growth arrest after damage (136). p53 sequence-specific DNA binding occurs through the core DNA-binding domain (137).

After stress-induced stabilization, p53 forms homotetramers that bind to two copies of a ten-base pair (bp) nucleotide sequence (5’-PuPuPuC(A/T)(T/A)GpyPyPy-3’) divided by a 0-13 bp spacer (138). The importance of p53 sequence-specific DNA binding is emphasized by the fact that most mutations found in human tumors within the p53 gene occur at “hot spots” within this DNA-binding domain (139). These mutations result in a dominant-negative form of the protein that can heterodimerize with wild-type p53 protein and disrupt its DNA binding, and thus inhibit downstream gene activation. The C- terminal tetramerization domain is (as its name implies) responsible for tetramer formation, as well as for regulation of DNA binding by the core domain (136).

As mentioned above, p53 has a very short half-life in unstressed cells, and thus cellular protein levels are strictly regulated at low levels by a variety of mechanisms. The stability and activity of p53 can be influenced in several ways: (1) through - mediated proteolytic degradation; (2) through subcellular localization of the p53 protein; and (3) through allosteric regulation of the core DNA-binding domain by the C-terminal regulatory domain. In the first instance, the Mdm2 protein, which acts as an E3 ubiquitin

29 ligase for p53 (140-142), binds to the N-terminus of p53 and results in its rapid degradation through the proteasome pathway. Mdm2 also blocks p53 activity by binding to the region of the peptide required for transcriptional activation of downstream genes. p53 is a transcriptional activator of the mdm2 gene and thus generates a negative feedback loop that tightly controls the amount of p53 present within the cell.

Phosphorylation of Ser15 and Ser20 within the Mdm2-binding domain of the p53 protein lowers the affinity of p53 for Mdm2 and inhibits degradation. Ser6, Ser9, Thr18, Ser33,

Ser37, and Ser46 are also phosphorylated within this domain and may enhance p53 stability as well as influence acetylation of the C-terminus (discussed below) (113, 122).

Additionally, phosphorylation of Mdm2 can also disrupt the interaction of these two proteins and result in stabilization of p53 (143).

For p53 to bind DNA and act as a transcription factor, the stabilized protein must be localized to the nucleus. In this fashion, activation of p53 can be regulated through its subcellular localization. p53 contains a nuclear localization sequence and a nuclear export sequence within its carboxy-terminal domain. Moll et al. (144, 145) demonstrated that in certain breast cancer and undifferentiated neuroblastoma cell lines, p53 is inactivated by sequestration in the cytoplasm. Mdm2 may also act to export p53 out of the nucleus (146), in addition to its ability to mediate p53 degradation.

The carboxy terminus of the p53 protein acts as an allosteric regulator of sequence-specific DNA binding. This was demonstrated initially by Hupp et al., (147) using a bacterially expressed protein. Recombinant bacterial p53 bound poorly to DNA, and binding could be enhanced by the addition of antibodies specific to the C-terminal region of the protein. Phosphorylation of Ser315 and Ser392 within this domain also

30 enhance sequence-specific DNA–binding. Dephosphorylation of Ser376 of p53 after IR allows the association of 14-3-3 proteins with the C terminus of the protein (148).

Stavridi et al. (149) demonstrate that this interaction is required for p53 to activate the

downstream gene, p21/waf1/cip1, and for the G1 cell cycle checkpoint arrest response.

Interestingly, this dephosphorylation event seems to be ATM-dependent, possibly by a phosphatase that is activated by ATM after IR (148). p53 also binds nonspecifically to short single-strand DNA fragments through its carboxy domain, and this binding may enhance sequence-specific DNA binding, thus providing a potential role for p53 in DNA repair after damage. However, the significance of p53 associating with single stranded

DNA is not known.

In addition to phosphorylation, the C-terminal domain of p53 can also be acetylated and sumolated in response to DNA damage. Acetylation and sumoylation both result in an increase in the transactivation ability of p53 and may account for this finding. In vivo, IR induces the acetylation of p53 at Lys320 by PCAF and Lys382 by

CBP/p300. Acetylation at these sites is dependent on N-terminal phosphorylation at Ser15 and to a lesser extent on phosphorylation at Ser6, Ser9 and Thr18 (122, 150). All of these phosphorylation events are ATM-dependent, although only Ser15 has been shown to be phosphorylated directly by ATM. Sumoylation occurs at Lys386 after DNA damage

(151). Sumoylation refers to the covalent attachment of a small ubiquitin-like molecule

(SUMO-1) to Lys residues, but in contrast to ubiquitination, does not result in proteolytic degradation. The significance of sumoylation after IR still remains to be elucidated.

31 p53 stabilization by IR

The signals upstream of p53 stabilization and activation after IR exposure most likely originate from DNA DSBs. This is supported by the fact that the kinases implicated in the phosphorylation of p53 are also implicated in DSB repair. These include two kinases that belong to the PI-3 kinase family, DNA-PK and ATM (see above), and indirectly, the checkpoint kinase 2 (Chk2). Ser15 of p53 was identified as a substrate for DNA-PK in vitro (152). Since DNA-PK is required for DSB repair in mammalian cells after IR exposure and since this residue falls within the Mdm2-binding domain of the protein, this was an attractive model of p53 stabilization after IR.

However, several recent reports have shown that Ser15 is phosphorylated in DNA-PK deficient cells, and that p53 accumulation in these cells is capable of generating cell cycle arrest or apoptosis after IR (153-155). ATM is another kinase that has been shown to phosphorylate Ser15 in response to DNA damage induced by IR in vitro (156, 157). In contrast to DNA-PK–deficient cells, AT lymphoblasts that lack the ATM gene have a decreased ability to generate p53-dependent cell cycle arrest or apoptosis after IR, suggesting an important role for ATM in generating these responses in vivo.

Additionally, ATM has been implicated in phosphorylating Ser9 and Ser46 directly and

Ser20 indirectly through Chk2 (158, 159). Interestingly, ATM has also been implicated as the kinase that phosphorylates Mdm2, thus disrupting the p53-Mdm2 interaction by phosphorylation of both of these proteins.

32 p53 target genes

As mentioned previously, the downstream target genes of p53 can be grouped generally into those required for growth arrest and those required for p53-dependent apoptosis (136). p21/waf1/cip1, 14-3-3σ and GADD45 are three genes that are activated by p53 after IR and that are involved in growth arrest. p53 directly transactivates the p21/waf1/cip1 gene, which is an inhibitor of cyclin-dependent kinases (CDKs). By

inhibiting CDKs, p21 blocks the cell cycle at the G1-S transition as well as the G2-mitosis transition. Additionally, p53 can help to block cells in the G2 transition through its activation of the 14-3-3σ gene, which sequesters proteins required for entry into mitosis in the cytoplasm. GADD45 interacts with the repair protein PCNA and inhibits the entry of cells into S phase. It is important to note that the studies identifying these genes as targets of p53 transactivation were done with high doses of IR. It has been shown that p21/waf1/cip1 and GADD45 are induced after low-dose IR exposure (>2 cGy) (160), but the exact role of p53 in cellular responses to low doses of IR are not known.

The p53-activated genes involved in IR-induced apoptosis are less clear and appear to vary with cell type. Bax was the first apoptotic gene identified that was transactivated by p53 and has been implicated in IR-induced apoptosis in testicular germ cell lines (161, 162). A recent paper by Fei et al. (163) describes the tissue-specific induction of several p53-dependent apoptotic genes (KILLER/DR5, Bid, Puma, Noxa) after 5 Gy total body irradiation. Other groups have also investigated the transactivation of apoptotic genes after high dose IR (163-165), but little is known about the p53- dependent activation of these genes after clinically relevant doses.

33 IR stimulation of NF-κB responses

NF-κB is a transcription factor that exists as a latent form in association with the inhibitory protein IκB. It was first discovered as a nuclear factor that bound to the κB element present in the immunoglobulin kappa light chain gene in 1986 (166). It soon became apparent that it exists in most mammalian cell types as an inducible complex sequestered in the cytoplasm [reviewed by (167, 168)]. Many signaling pathways are known to cause degradation of IκB to release free NF-κB that then migrates into the nucleus and regulates induction of a wide variety of genes through decameric κB-binding sites. NF-κB target genes are known to play roles in a diverse array of cellular and physiological functions, including apoptosis, proliferation, cell adhesion, migration, and inflammatory and adaptive immune responses (reviewed in (169-172). While most studies focusing on the activation mechanisms and functional significance of NF-κB involve the use of pro-inflammatory cytokines, such as -α (TNFα) and IL-1, NF-κB has also been known to be activated by IR since 1991 (126). Using electrophoretic mobility shift assays (EMSAs) to detect newly liberated nuclear NF-κB, the above group showed that an IR dose as low as 2 Gy was sufficient to activate NF-κB in the KG-1 human myeloid cell line, with the peak dose at 5 to 20 Gy and the peak time around 2 to 4 hours after irradiation (126). Like in many other NF-κB activation pathways known, a protein synthesis inhibitor did not prevent this activation pathway, indicating that de novo protein synthesis is not required. Since the above report, NF-κB activation by IR has also been observed in different cell systems (see below).

34 The IR doses required for maximal NF-κB activation vary greatly depending on the cell lines or systems analyzed. In the EBV-transformed 244B human lymphoblastoid cell line, as in the case for the activation of AP-1 (see below), 0.5 Gy of IR was shown to yield the maximal NF-κB activation (127, 173). In the U1-Mel human melanoma cell line, 3 to 4.5 Gy was optimum with higher doses giving reduced activation (127, 128).

By contrast, in human diploid fibroblasts, even doses as high as 20 Gy failed to activate

NF-κB (174). More complicating are the contrasting observations of NF-κB activation following whole body irradiation at different IR doses. Zhou et al. (175) reported that

NF-κB activation by IR (8.5 Gy) is tissue-specific and only detectable in bone marrow, lymph nodes, and spleen. In contrast, Li et al. (176) demonstrated the activation of NF-

κB in liver and when mice were exposed to 20 Gy of IR. In cerebral cortex of

Sprague-Dawley rats, NF-κB activation was detectable at 5 Gy with a steady increase up to 30 Gy (177). It thus appears that different normal tissues possess differential IR sensitivity toward the activation of NF-κB in vivo. However, it is presently unknown whether lower, more clinically relevant doses of IR (<2 Gy) would also yield activation of NF-κB in these normal tissues. Since NF-κB activation by IR can be readily observed in different human cancer cell lines derived from various tissue types in vitro (see below), but not normal counterparts (S. Miyamoto, unpublished observations), it is intriguing to consider the possibility that transformation or carcinogenic events may sensitize cells to

IR for activation of NF-κB. Ashburner et al. (174) found that immortalization of human fibroblasts by SV40 was correlated with acquisition of NF-κB activation capacity following IR treatment, further supporting this idea. Thus, it is possible, and even likely,

35 that cancer cells in vivo may display higher sensitivity to IR for activation of NF-κB, when compared to normal counterparts.

While activation of NF-κB can be affected by tissue types and transformed states, in those systems where NF-κB activation by IR can be readily observed distinct molecular components and signaling events have been implicated under different experimental settings. These include protein tyrosine kinases (PTKs), PKC, ROS, Ras,

ATM, DNA-PK, IκB kinase (IKK), and the ubiquitin-proteasome pathway (Figure 1.4).

Of particular interest with regard to the mechanism of activation of NF-κB by IR is the possibility that nuclear DNA damage may be directly involved in the activation of the cytoplasmic NF-κB/IκB complex, i.e., a novel nuclear-to-cytoplasmic, retrograde signaling pathway. Huang et al. (178) demonstrated multiple lines of evidence for the existence of such a signaling pathway induced by another anticancer agent, camptothecin.

IR is known to cause DSBs and subsequently activate nuclear kinases, such as ATM

[reviewed in (179)]. ATM was critical for NF-κB activation by IR at 20 Gy using an

ATM-/- mouse model, and the kinase inhibitor wortmannin prevented such activation

(180). However, the requirement of ATM in NF-κB activation seems to be different at different IR doses. Earlier work by Jung et al. (181, 182) showed that NF-κB activation in the SV40-immortalized AT5BIVA human fibroblast cell line was still observed in these ATM-deficient cells exposed at 6 Gy of IR. However, the same group reported that

ATM was required for Iκ B degradation, nuclear translocation of NF-κB and transcriptional activation by NF-κB in the same cell systems upon exposure to 20 Gy of

IR (183). Thus, the contribution of ATM is apparently dependent on the dose of IR in this cell system. Moreover, the general role of ATM has been challenged by Ashburner

36 et al, (174), since neither AT nor wild-type human diploid fibroblasts (not immortalized by SV40) activated NF-κB. Similarly, while another nuclear ATM-related kinase, DNA-

PK, has been implicated in the activation of NF-κB after 10 Gy exposure of HeLa cervical carcinoma and M059K glioma cell lines (184), liver and kidney tissues from

DNA-PKcs, Ku70 and Ku80 knockout mice displayed no NF-κB activation defects when exposed to IR at 20 Gy (185). In the latter report, c-ABL was also found dispensable for this activation pathway at this high dose of radiation. Moreover, there are also conflicting results regarding the role of the IκB kinase complex, phosphorylation and degradation of IκB by the ubiquitin-proteasome pathway to release active NF-κB when cells are exposed to different doses of IR in different cell systems (185-190). While direct phosphorylation of IκB by both ATM and DNA-PK has been shown in vitro (182-

184), whether such a direct phosphorylation is indeed induced by IR in intact cells has not been established. Thus, it is presently unclear whether nuclear DSBs are indeed required for activation of NF-κB by IR, especially at clinically relevant doses.

A previous study mentioned above (126) implicated the involvement of PKC based on the effect seen with H7 inhibitor and down regulation of PKC by prolonged

TPA treatments. Using the human Ramos B cell line exposed to IR doses of 10 to 40 Gy,

Uckun et al. (191) have implicated the involvement of PTKs and possibly PKC in the activation of NF-κB. However, by comparing the kinetics and magnitudes of NF-κB responses in primary tonsillar B cells exposed to TPA versus 15 Gy of IR, Wilson et al.

(192) concluded that simple PKC activation is unlikely to account for this activation process. Uckun et al. (191) also suggested that the generation of ROSs might be upstream of PTK activation. ROSs were previously proposed to represent a common

37 second messenger for NF-κB activation by distinct signaling agents (193). Consistently, antioxidants (e.g., N-acetylcysteine (NAC)) have been shown to prevent NF-κB activation by IR in several cell lines exposed to a dose range of IR (0.5-30 Gy) (194-196).

However, accumulating evidence indicates that the role of ROSs in NF- κB activation is also cell-type–restricted (reviewed by (197-200)), and NAC can directly inhibit 26S proteasome activity that is essential for degradation of IκB to release free NF-

κB (201). Moreover, the site of IκBα degradation induced by IR can also be distinct from those induced by cytokines, since it was suggested to occur at the plasma membrane, rather than in the cytosol, in U251 glioblastoma cell line exposed at 10 Gy (202). The same group also indicated the involvement of the Ras signaling pathway in NF-κB activation by IR through the use of a recombinant adenoviral system that expresses the anti-Ras single-chain antibody fragment (190). Thus, there is the possibility that the activation of NF-κB may occur at the level of the plasma membrane at high doses of IR exposure. The upstream regulators or how the Ras pathway connects with NF-κB activation following IR exposure is not well defined yet. It is also unclear whether lower doses (<2 Gy) of IR also employ the same activation mechanism to release NF-κB in the above or other cell systems.

Putting these published observations together, it is apparent that there is no consensus on the activation mechanism for IR regulation of NF-κB activity, especially at clinically relevant doses of IR. It is unclear whether IR-induced DSBs represent the signal initiating events or whether other stresses (e.g., ROSs) mediate this activation pathway. Different IR doses seem to invoke distinct signaling events to cause activation

38 Figure 1.4. IR stimulates the phosphorylation of IKK, which results in the degradation of IκB allowing for NF-κB to move into the nucleus and activate downstream genes.

39 of NF-κB. More complicating is the undetectable nature of NF-κB activation in many normal tissues after whole body irradiation, even at a relatively high dose of IR (up to 8.5

Gy). Many diploid human, normal cell types are also refractory to NF-κB activation following IR treatment in vitro, even though TNFα can efficiently activate NF-κB in these cells. These observations contrast with findings that many human cancer cell lines derived from different tissue types are permissive for activation of NF-κB by IR exposure in vitro. It is, however, important to note that NF-κB activation was measured by

EMSAs in most of these studies. Recent studies indicate that NF-κB activation, as measured by reporter gene activation, can take place without an apparent increase in the

NF-κB DNA-binding activity (203). Further in-depth analyses of signaling pathways and differential IR sensitivity of human normal versus cancerous cell types for activation of

NF-κB are, therefore, warranted.

What is the role of NF-κB activation in biological response to IR? An earlier study showed that induction of IL-6 by 5 Gy of IR exposure of FH109 human embryonic fibroblast cell line was dependent on the κB-binding site present in its promoter (126).

Others found a similar observation in PC-12 rat pheochromocytoma (204) and HeLa cell lines (205). IL-6 is critical for acute inflammatory reactions seen in IR-treated tissues.

TNFα, IL-1α, IL-1β, intercellular adhesion molecule-1 (ICAM-1) and E-selectin are also critical for inflammatory responses. Zhou et al. (206) found that induction of TNFα, IL-

1α, IL-1β and IL-6 in spleen, mesenteric lymph nodes and bone marrow of Balb/c mice following exposure to 8.5 Gy of IR was significantly reduced in mice knocked out for the p50 subunit of NF-κB. Studies also indicated that ICAM-1 induction by IR in HeLa,

HaCaT keratinocyte line, human umbilical vein endothelial cells (HUVECs) and

40 microvascular endothelial cells depends on NF-κB activation (195, 207). Similarly, E- selectin induction by IR (0.5-10 Gy) was also found to depend on activation of NF-κB in

HUVECs (208). All of these genes possess κB sites in their promoters (reviewed in

(171)). Thus, there is a consensus that the synthesis of inflammatory cytokines and molecules following IR exposure is coordinated by the activation of NF-κB in both in vitro and in vivo situations.

Another important role of NF-κB in biological response to IR is the modulation of radiosensitivity. An earlier study showed that the radiosensitive phenotype of SV40- immportalized AT5BIVA cells exposed to 6 Gy of IR was reversed by the expression of an N-terminally truncated I-κB that inhibited constitutive NF-κB activity present in these cells (181). These studies suggested that NF-κB promoted radiosensitivity. The similar radiosensitizing role of NF-κB has been observed in the human MRCA5CV1 fibroblast line exposed to 5 Gy (209) and U1-Mel cells exposed at the dose range of 5 to 15 Gy

(128). In contrast, Wang et al. (210) showed that when the HT1080 fibrosarcoma cell line was treated with IR, inhibition of NF-κB activation by the stable expression of the

“super-repressor” IκB mutant caused significant radiosensitization, as measured by colony forming ability assays at doses of 5 and 10 Gy (but not at 2.5 Gy). This latter study indicated that NF-κB activation caused radioresistance, rather than radiosensitization. This finding has also been confirmed in the LOVO colorectal cancer cell line in vitro and in tumor volume reduction of its xenograft in vivo (211). Similarly, the role of NF-κB in radioresistance was also seen in the A172, MO54 and T94 human glioma cell lines at 1 to 10 Gy (212, 213), the 3SB mouse lymphoma cell line exposed to

5 Gy (214), the UM-SCC-9 head and neck carcinoma cell line at a dose range of 1 to 12

41 Gy (215), the KB head and neck carcinoma cell line at 2 Gy (216), HeLa cells exposed to

2 Gy (217), and the HK18 human keratinocyte line at 2 to 12 Gy (218). These observations indicate that radiosensitization can be achieved by inhibiting NF-κB activation at the clinically relevant IR doses in many cancer cell types. Like the unresolved signaling mechanisms discussed above, the role of NF-κB in radioresistance is not straightforward and is heavily dependent on the specific cancer cell lines examined.

Several studies found that alteration of NF-κB activation did not affect radiosensitivity, either positively or negatively. These include the PC3 prostate cancer and HD-MyZ

Hodgkin’s lymphoma cell lines (219, 220), the U251 glioblastoma cell line (190, 202) and the RIE-1 rat intestinal epithelial cell line (221). Russell et al., (190, 202) further implicated the lack of NF-κB–dependent radioresistance in SF539 glioblastoma, SW620 and HT29 colon carcinoma, BXPC-3 pancreatic carcinoma, and C-39A normal fibroblast cell lines.

In conclusion, activation of NF-κB by IR can be observed in different cell systems at clinically relevant doses; however, there are also cell contexts in which it is not readily observed by EMSAs. In particular, activation of NF-κB in normal tissues versus primary tumors in vivo at low-dose IR needs further evaluation. Use of different assays to detect

NF-κB activation may also be required to fully appreciate the potential tissue specificity of this activation pathway in normal cells. In this context, a model developed by Carlsen et al. (222) to directly image NF-κB activation in vivo or development of similar assay systems could be most useful. While NF-κB activation by IR can be readily observed in a wide variety of cancer cell lines in vitro, the signal transduction mechanisms are complex and likely distinct at low versus high doses of radiation. There seems to be a

42 consensus regarding the role of NF-κB in mediating inflammatory responses to IR exposures, but its role in radiosensitization is highly dependent on the cancer cell line.

Importantly, it is entirely unclear whether these in vitro observations regarding the mechanisms and roles of NF-κB activation can be extrapolated into primary and metastatic tumors in cancer patients. Other studies employing 2D- versus 3D-culture systems have demonstrated that the sensitivity of specific cancer cell lines to radiation and anticancer drugs can be drastically altered by the presence of specific extracellular matrix components as well as cell polarity (223, 224). More importantly, activation of

NF-κB by etoposide and TNFα was also greatly modulated by the 3D architecture of the

T4-2 mammary tumor cell line (224). Since most of the aforementioned studies were performed under 2D conditions on plastic, it is critical to determine both the mechanisms and the roles of NF-κB activation by IR in modulation of radiosensitization under 3D culture and in vivo conditions. Even with these theoretical shortcomings, data to date are promising that NF-κB represents a novel anticancer target to enhance radiotherapy at clinically relevant doses. Such efforts have been spearheaded by multiple investigators, especially Baldwin and his colleagues (211, 225-227), to enhance not only radiotherapy but also . One of the major questions that remain to be resolved is: What are definable mechanism(s) that explain whether NF-κB plays a positive, negative or neutral role in modulation of radiosensitivity in specific cancer cell types in vivo?

Answers to this question will undoubtedly help target specific cancer types for enhancement of radiotherapy through NF-κB modulation, but will require analyses of

NF-κB activation mechanisms at clinically relevant doses of IR.

43 AP-1 transcription factor activation by IR

The ‘activating protein-1’ (AP-1) transcription factor was first identified as a protein responsible for regulating the expression of the human metallothionein promoter

(228, 229). Using AP-1 DNA-binding–site sequence-specific chromatography, several proteins were purified suggesting that the AP-1 transcription factor was a complex of more than one protein. Subsequently, analyses identified these proteins as members of the c-jun and c-fos protein families that can homo- or heterodimerize. The functional

AP-1 transcription activator complex was subsequently identified as containing the c-Jun,

JunB, JunD, as well as c-Fos, FosB, Fra-1 and Fra-2 proteins (230-235). Further studies identified additional binding partners (co-activators or co-repressors), such as ATF-2

(activating transcription factor-2) (236-240), Maf (241, 242), Nrf1 (238) and Nrl (242).

Jun proteins and their regulation

The v-jun oncogene was first isolated from the avian sarcoma virus-17 (jun is derived from ju-nana, the Japanese word for 17) (243). c-Jun and a number of Jun family members were isolated from mammalian cells based on their abilities to homo- and/or heterodimerize with themselves, Fos family members, or CREB/ATF, Maf and Nrl.

Various compositions of AP-1 result in differential DNA-binding affinities (e.g., c-Jun/c-

Fos > c-Jun/c-Jun > JunD/JunD > JunB/JunB), and some of these complexes can interfere with other AP-1 complexes (244, 245). It was shown, for example, that JunB may act as either a positive or negative regulator of transcription via binding to an AP-1 site (197,

246, 247). c-Jun protein is activated by phosphorylation of its N-terminal domain on residues Ser63 and Ser73. Substitution of these serine residues with alanine completely

44 inactivated Jun-dependent transcription (248). Analyses of c-Jun activation following UV irradiation led to the cloning of c-Jun N-terminal kinase (JNK), a member of the stress- activated protein kinase (SAPK) family (249).

Fos proteins and their regulation

The Fos protein family derived its name from the Finkel-Biskis-Jenkins murine osteosarcoma virus, from which v-Fos was isolated. c-Fos is a 55 kDa nuclear phosphoprotein (250). Fos family members are expressed predominantly in osteoblasts. c-Fos forms exclusive heterodimers with its binding partners, including the Jun family members ATF, Maf and Nrl, which are unable to form homodimers (242, 251). As in the case of c-jun, c-fos transcriptional activity is thought to be regulated by phosphorylation.

An enzyme called Fos kinase was isolated (252, 253), however, its role in the activation of AP-1 by IR has yet to be determined.

AP-1 characteristics

A common feature of proteins that form the AP-1 transcription factor complex is the presence of a leucine zipper that consists of heptad repeats of leucine residues aligned along one face of an alpha helix. Intercalation of leucine residues within the zippers of the two proteins and formation of a coiled-coil structure result in an AP-1 protein dimer.

Basic regions adjacent to the leucine zippers serve as the DNA-binding domain of the

AP-1 factor. The AP-1 binding site, also known as the TRE (12-O-tetradecanoylphorbol- b-acetate (TPA)-response element) (229), has a palindrome consensus sequence

TGA(C/G)TCA. The AP-1 site is similar to the cAMP response element (CRE):

45 TGACGTCA. Binding of the Jun family members to the CREs either alone or in complex with ATF/CREB proteins was reported (245, 254). Many genes have AP-1 site(s) in their promoters; however, these are mostly genes that are relevant to bone biology (e.g., collagenase) and that are responsive to AP-1 induction (255). The induction of IL-6 by IR (2 Gy, 6 h) in lymphocytic cancer cell lines was shown to be at least partially dependent on AP-1 activity (126, 256)

AP-1 activation after IR

Soon after AP-1 was discovered, elevations in the steady state levels of c-jun, junB and c-fos mRNA after IR were reported (131, 257, 258). c-fos mRNA was induced in syrian hamster embryo cells within 3 h after 0.75 Gy of x-rays or 0.9 Gy of gamma- rays (258). In contrast, no detectable c-fos mRNA induction was observed after treatment with high-LET fission-spectrum neutrons. In separate experiments, increases in c-jun, c- fos and junB steady-state mRNA levels in human HL-60 human promyelocytic leukemia cells were not detected at IR doses below 5 Gy (131). fosB and junD were also shown to increase their mRNA steady-state levels after ≥5 Gy of IR in HL-60 cells (259).

Curiously, the majority of studies of induction of AP-1 components by IR have been performed on cells not related to bone tissue, where these proteins, particularly c-Fos, are supposed to have physiological significance. Furthermore, the doses of IR used were much higher than those commonly used in radiotherapy. After supra-lethal doses of IR, increases in c-jun and c-fos mRNA levels correlated with a number of detectable intracellular processes, such as transient down regulation of cdc2, cyclin A, cyclin B, and cdc25 genes (260), induction of interleukin-1 (261), and apoptotic DNA fragmentation

46 (262). However, no direct connection between these events was shown, and no transcriptional activation of AP-1 nor direct increases of c-Fos, c-Jun or AP-1 proteins were determined.

Responses of AP-1 components to the lower, clinically relevant doses of IR (0.25-

2 Gy) and induction of AP-1 activity as a transcription factor were later shown in

Epstein-Barr virus- transformed human lymphoblastoid 244B cells. Induction of c-fos mRNA, after as low as 0.25 Gy, and c-jun, after 0.5 Gy, was observed, and levels of these mRNAs peaked at 1 h, but decreased 4 h after IR treatment (263). Subsequent studies of the c-jun promoter region, using a chloramphenicol acetyl transferase (CAT) reporter, revealed that AP-1 and CCAAT DNA elements were required for induction of this gene by high-dose IR exposures (264). These data suggested that a positive feedback loop was possible in c-jun activation following supra-lethal doses of IR. The requirement of AP-1 for the induction of jun gene was further confirmed in hypoxic HeLa cells (265). These data suggested the activation of a pre-existing AP-1 complex by IR via a post- translational mechanism(s). Indeed, the purification and cloning of JNK led to the elucidation of the c-Jun phosphorylation mechanism and AP-1 induction following various stresses, including IR (248, 249). Studies linking JNK activation and increased

AP-1 activity after low doses of IR remain to be performed in detail.

Subsequent studies have suggested that induction of c-Jun was dependent on PKC activation. This conclusion is based on evidence that depletion of PKC by pretreatment with TPA inhibited post-IR induction of the transcription of AP-1 components c-jun and c-fos (266). Also, TPA was shown to induce JNK activity (267). A similar dependence on PKC was demonstrated for c-fos gene induction after IR (268). New evidence derived

47 from the study of ataxia telangiectasia (AT) cells indicated the role of ATM in post-IR c-

Jun phosphorylation and activation (269). JNK is phosphorylated and activated by the

SAPK/ERK kinase (SEK1) pathway, also known as MKK4. MKK4 is, in turn, phosphorylated by MEKK1. Activation of MEKK1 after stress could be the result of activation of p21-activated kinase (PAK) via the Rho subfamily of small GTPases, in particular Rac1 and Cdc42. In fact, dominant-negative mutants of Rac1 or Cdc42 inhibited JNK activation stimulated by growth factors [reviewed in (270)]. It is not clear, however, whether these small GTPases participate in the IR induction of c-Jun phosphorylation. It is also yet to be determined if AP-1 activation and induction of c-fos or c-jun transcription via high-dose IR stimulation of JNK activity occurs after low, clinically relevant doses of IR. To date, nearly all of the studies concerned with elucidation of the signal transduction mechanisms of activation of the AP-1 transcription factor have used supra-lethal doses of IR.

Even though activation of AP-1 after IR has been described, it is not clear that activation of this transcription factor mediates downstream genes that have functional consequences for cellular responses to IR. Among the genes that have been clearly shown to be activated by AP-1 after IR are various cytokines, including TGF-ß1 (271), vascular endothelial growth factor VEGF (272), and IL-6 (205). Interestingly, none of these genes have been found to be responsive to AP-1 activation alone. In the case of

TGF-ß, it was shown that AP-1 activation alone was, in fact, not sufficient for the induction of the gene transcript (273). Likewise, AP-1 acts synergistically with NF-κB

(205) to induce IL-6 transcription. Clearly, cells have evolved complex mechanisms for gene induction after IR, as well as other stresses. This fact is highlighted by the failure of

48 multiple researchers to find a majority of known AP-1 transcriptionally dependent downstream genes induced in cells under conditions known to activate AP-1 after IR.

Additionally, it should also be noted that AP-1 activation was not found in all cells after

IR exposure (127), indicating cell specificity in this IR-responsive transcription factor and probably in JNK/SAPK/MEKK1 activation.

Sp1 activation after low doses of IR

Sp1 was the first identified member of a family of transcription factors composed of sixteen different proteins that bind to similar sequences in the promoter regions of human genes (274). Sp1 is ubiquitously expressed in mammalian cells and binds with high affinity to GC-rich sequences called GC-boxes, and with a lower affinity to CACC- elements called GT-boxes. Sp1 is thought to play a role in various cellular processes, including cell cycle regulation, chromatin remodeling and maintenance/propagation of methylation-free CpG islands. The importance of Sp1-mediated transcription is underscored by the fact that Sp1 null mice are embryonic lethal by day 10. This family of transcription factors includes the Sp family members (Sp1, Sp2, Sp3 and Sp4) and the kruppel-like family members. Sp1 and Sp3 have additional isoforms as a result of trans- splicing between pre-mRNAs or alternative translation initiation, respectively, that add to the complexity of this family of transcription factors. Additionally, Sp3 can act to repress transcription of Sp1-dependent promoters by competing with Sp1 for the DNA-binding site.

Sp1 and Sp3 are retinoblastoma (Rb) control proteins (RCPs) that bind to promoter regions, known as retinoblastoma control elements (RCEs). RCPs are regulated by the Rb protein and control induction of several immediate early growth-response

49 genes (IEGs) after cell stress (i.e. c-fos, c-jun, c-myc and TGF-β1). Two reports have shown the inducibility of Sp1 DNA binding after high doses of IR (>4 Gy). Activation of

Sp1 DNA binding after lower doses of IR has not been investigated in detail to date, although changes in Sp1 post-translational modification were observed after as little as 2 cGy IR (94). Yang et al. (128) demonstrated increased Sp1-related, RCP DNA binding to an RCE in radioresistant human malignant melanoma (U1-Mel) cells after 4.5 Gy, which correlated with an increased expression of IEGs; although low doses of IR were, in general, not used in this study, exposure of U1-Mel cells to this dose of IR resulted in only 60% lethality and greater than 50% survival within the 4-h potentially lethal damage recovery period allowed. Meighan-Mantha et al. (132) demonstrated a 5-fold increase in

Sp1-binding activity in head and neck squamous cell carcinoma after 15 Gy. Further research needs to be done to explore the possible role(s) of Sp1 and the complex RCP binding to RCEs within specific genes during cell death or survival responses after low- dose IR exposures.

IR-activation of the Egr-1 (Sp1-related) transcription factor.

The early growth response 1 gene (Egr-1, a.k.a., NFGI-A, Krox-24, TIS-8, or zif/268) product is unique among transcription factors, since both the transcription factor

(129, 275) and the gene transcript (129, 276) are both activated and induced, respectively, by supra-lethal doses of IR. The Egr-1 gene encodes a nuclear phosphoprotein with a cysteine-histidine zinc finger motif that is homologous with the Wilm’s tumor susceptibility gene and binds to the Sp1-like DNA element 5’-CGCCCCCGC-3’ (277,

278). Egr-1 belongs to a class of transcription factors that includes Sp1 (278), TFIIIA

50 (279), and SW15 (280). Along with the reported IR inducibility of Egr-1, the gene is also reported to be induced by various mitogenic stimuli, all believed to be the result of v-Src and v-Fps protein kinase activities (281, 282).

To date, the minimum dose of IR required to induce Egr-1 transcription or activation of transcription factor binding is 5 Gy, and has been reported in a variety of cancer cell lines, including human HL-525 myeloid leukemia, U87 malignant glioma, and

A375-C6 melanoma as well as in the AsPc-1 human pancreatic tumor cells (129, 260,

283-285). Induction of Egr-1 was prevented by the free radical scavenger NAC (129), however, the transcription factors responsible for regulating the gene’s promoter at

proposed serum response elements (SREs) and/or CC(A/T)6GG sequences remain unexplored (129, 283, 284, 286). Only one study has investigated the potential functional significance of Egr-1 induction on radiosensitivity, specifically using the A375-C6 melanoma cell line (284). Using stable expression of an Egr-1 dominant-negative mutant, as well as transient expression of Egr-1 antisense, Egr-1 levels were diminished and the effects of these changes on radiosensitization in terms of ‘percent growth inhibition’ were monitored following IR (5 and 20 Gy doses were used). Rather modest effects of antisense and dominant-negative Egr-1 expression on radiosensitivity were reported. The results were interpreted to indicate that Egr-1 is required for the growth- inhibitory response of these melanoma cells to IR. Egr-1 knockout studies and the use of clinically relevant doses of IR using colony forming ability assays would be required for definitive proof of this hypothesis.

Considering that the Egr-1 promoter has been one of the few IR-inducible sequences selected for ‘proof of principle’ gene therapy (discussed below) for the

51 proposed delivery of the bystander death gene, TNF-α (192, 208, 287, 288), it is surprising that: (a) such little information is available about the signal transduction pathways that regulate Egr-1 induction in a variety of cell types; and (b) there is paucity of data to indicate that the Egr-1 promoter can be induced by doses of IR that would be considered relevant to clinical therapy. Analyses of the published literature revealed that only one study has examined the signaling pathways that may stimulate the Egr-1 promoter at somewhat reasonable doses of IR (~5 Gy) (286). In this study, the Erk1/2 and SAPK/JNK signal transduction pathways were implicated, based entirely on the use of inhibitors of these specific pathways. The role of PKC was also implicated (130, 289-

291), however, these studies utilized supra-lethal doses of IR to leukemic cells, and the reported results were not applicable to clinically relevant doses of IR. A further review of the literature indicates that not all studies have found Egr-1 induction in mammalian cells after IR, and that in some cells the induction or expression of Sp1 can interfere with specific regulatory elements in the Egr-1 promoter (275). In summary, stimulation of

Egr-1 is observed in human cancer cells, however, supra-lethal doses of IR are generally required for the observed changes in transcription factor activation and induction.

Oct-1 and NF-Y transcription factors

A number of recent papers have demonstrated activation of octomer binding protein 1 (Oct-1) and/or NF-Y transcription factors in response to stress (132, 292-294).

Most of the data reported to date have used agents other than IR, such as methylmethane sulfonate (MMS), camptothecin, etoposide, cisplatin, or ultraviolet radiation (UV). A significant induction of Oct-1 (5-fold), monitored by electrophoretic mobility assays

52 (EMSAs), was observed in PCI-04A head and neck squamous cell carcinoma cells within

4 h after 5 Gy (132). Similar, albeit significantly less, Oct-1 binding activity was also reported in PC-3 human prostate cancer, as well as MDA-MB-231 human breast cancer cells after 5 Gy. Activated Oct-1 binding was more rapid and prolonged (lasting 0.5 –

16 h post-treatment) using supra-lethal doses of most cytotoxic agents; for example, such responses were reported in PCI-04A cells after 20 Gy. Lower doses of IR have not been explored, and changes in promoter activity before vs. after IR, using an Oct-1-dependent promoter, remains to be performed.

As with many activated transcription factors, particularly after exposure of cells to supra-lethal doses of IR, it is unclear as to the exact function of activated Oct-1 and/or

NF-Y DNA-binding activity on gene expression. One study found simultaneous activation of Oct-1 and NF-YA DNA-binding activities were essential for the p53- independent induction of the GADD45 promoter after lethal doses of MMS (100 µg/ml) or UV (10 J/M2) in HCT116 human colon carcinoma cancer cells (295). Separate mutations within the two Oct-1–binding sites or the NF-YA CAAT box abrogated induction of the GADD45 promoter. Unfortunately, IR exposures were not examined in this study. In a study of the secretion of von Willebrand factor (VWF) after IR, activated

NF-Y DNA-binding activity to a CCAAT element within the –90 to +22 core promoter was found to be essential for this gene’s promoter activity after lethal doses of IR (294).

Since VWF is an important factor in thrombus formation, understanding the regulation of this secreted factor after IR may be very important for controlling IR-induced thrombus formation.

53 The activation of Oct-1 DNA binding has not been a universal property of IR- exposed cells. Sahjidak et al. (127) found that Oct-1 levels were not altered in radioresistant human melanoma cells before or after various doses of IR (2-10 Gy) compared with normal basal level DNA-binding activities using EMSAs and 50-bp oligomers containing the octamer-1–binding motif. Factors such as cell type specificity or specific growth conditions that may affect the activation of Oct-1 or NF-Y DNA- binding activities after IR in mammalian cells have yet to be determined. Analyses in vivo of VWF and Oct-1 or NF-Y activities may also be warranted.

Exploiting IR-inducible Promoters and Transcription factors for improved cancer therapy

Over the past ten years or more, an extensive effort by many laboratories has attempted to understand the cellular and molecular biology of stress induced by IR.

Unfortunately, most of these studies have focused on signal transduction events occurring after relatively high, non-clinically relevant doses of IR that subsequently activate transcription factors and corresponding downstream promoters (reviewed above, Table

1). The assumption has been that one can then extrapolate down to low doses of IR and assume that the same changes in signal transduction events, activation of transcription factors (as well as co-activators and co-repressors), stimulation of specific promoters, and induction of IR-inducible gene products occur at low, clinically relevant doses of IR. This strategy has been used in an attempt to exploit the Egr-1 promoter as described above

(290). To date, little success in this clever approach has been achieved. Overall, the accumulated literature on IR-inducible gene expression seems to indicate that a better

54 understanding of low-dose IR-inducible promoters of specific genes in solid tumors, and not the more radiosensitive lymphocytes, is in order. More research focusing on gene induction events occurring in solid tumors (or cells from solid tumors) after IR doses between 0.5-2 Gy is needed. Extrapolation from high (many times supra-lethal) doses of

IR to clinically relevant exposures may often be misleading.

Examination of the literature indicates that two transcription factors, p53 and NF-

κB, are reproducibly responsive (120, 296-303) to clinically relevant (and sometimes lower) doses of IR. These data suggest that promoters regulated by these transcription factors (e.g., the p21- or 3X-κB–containing promoters) may be useful for IR-inducible expression in gene therapy regimens. For example, Ueda et al. (304) used the NF-

κB–responsive c-IAP2 promoter to drive the expression of the pro-apoptotic Bax gene to induce cell death at 2 Gy. More research is needed to explore IR-inducible RCPs and

RCE-directed promoters. Since these transcription factors are activated under conditions where cells survive and recover after IR, more extensive analyses of their induction and signal transduction pathways are warranted.

55 Chapter 2: Repression of IR-inducible clusterin expression by the p53 tumor

suppressor protein

This work was published in Cancer Biology and Therapy

Cancer Biol Ther. 2003 July 1; 2:4; 372-380

ABSTRACT

The clusterin (CLU) protein has been reported to have both cytoprotective and cytotoxic activities. Previous data from our lab suggest that the secretory form of CLU

(sCLU) is cytoprotective and induced after very low, nontoxic doses of ionizing radiation

(IR: >0.02 Gy), while a nuclear form is cytotoxic (95). Cells must presumably suppress sCLU to stimulate cell death, however, factors regulating the stress-inducible expression of sCLU have not been elucidated. Here we demonstrate that p53 can suppress sCLU induction responses. A variety of cytotoxic agents stimulated sCLU expression and DNA damage was sufficient but not necessary for induction. IR-stimulated CLU promoter activity, with concomitant increases in CLU mRNA and protein, showed that CLU induction was delayed with maximal expression observed 48-96 h post-treatment.

Expression of the human papillomavirus E6 protein in MCF-7 breast or RKO colon cancer cells enhanced basal sCLU levels. Isogenically matched HCT116 colon cancer cell lines that differed only in p53 or p21 status, confirmed a role for p53 in the transcriptional repression of sCLU. Loss of functional p53 in HCT116:p53-/- cells augmented sCLU de novo synthesis after IR exposure. Repression of sCLU protein

56 levels by p53 may be important for the cascade of p53-mediated events leading to cell death after IR or other cytotoxic agent exposure.

INTRODUCTION

Clusterin (CLU) is a sulfated glycoprotein implicated in many physiological and pathological processes, including tissue remodeling (21), complement inhibition (14,

305), lipid transport (9, 37), multiple sclerosis (306), atherosclerosis (307, 308) and

Alzheimer’s disease (39, 41, 309). Two different forms of the CLU protein exist: an 80 kDa secretory form composed of 40 kDa alpha and beta subunits and an ~55 kDa nuclear form. Our laboratory identified CLU as a x-ray inducible protein/transcript (xip8) (11).

We showed that a nuclear form of CLU (nCLU) associated with the DNA double strand break (DSB) repair protein, Ku70, and was a pro-death factor (35, 71). However, the secretory form of CLU (sCLU) did not associate with Ku70, and this form of CLU was induced by much lower, nontoxic doses of IR. In fact, sCLU was induced at ~0.02 Gy, a dose found to be growth-stimulatory and cytoprotective in many human cancer cells (71).

The regulation of sCLU remains poorly understood.

Elevated levels of sCLU protein and mRNA were noted in several different types of human malignancies (49, 57), and forced over-expression of sCLU in transformed cell lines resulted in an increased resistance to various chemotherapeutic agents (48, 62). In addition, abrogation of CLU mRNA expression following antisense expression lead to modest chemo- and IR-sensitizations in various cell lines (72-74, 310). These data support a cytoprotective role for sCLU.

57 The p53 tumor suppressor gene is mutated in over half of all human tumors (311).

Wild-type p53 protein is stabilized after cellular stress and acts as a transcription factor for various downstream genes, including Bax, p21 and GADD45, resulting in either cell cycle arrest or apoptosis (312-314). p53 can also act as a repressor of transcription, although exact mechanisms of transcriptional suppression still remain to be elucidated.

Examples of p53-repressed genes include presenillin (315), (316), cyclins A(317) and B (318) and cdc2 (319).

Our laboratory identified CLU as a x-ray inducible protein/transcript (xip8) (11).

We showed that a nuclear form of CLU (nCLU) associated with the DNA double strand break (DSB) repair protein, Ku70 (35, 71). However, sCLU did not associate with Ku70, and this form of CLU was induced by much lower, nontoxic doses of IR. In fact, sCLU was induced at ~0.02 Gy, a dose found to be growth-stimulatory and cytoprotective in many human cancer cells (71).

Although regulation of CLU gene expression following and testosterone exposures has been investigated (64, 320), the regulatory control of sCLU synthesis after

IR or other cytotoxic agents has not been elucidated. We show that CLU mRNA and protein synthesis in human cells is induced after various cytotoxic stresses, including exposure to many anti-tumor agents. IR-induction studies of CLU promoter activity, CLU mRNA accumulation, and sCLU protein synthesis confirm that sCLU expression occurs in a delayed fashion, with initial IR-activation of the CLU promoter occurring 24 h post- exposure, and mRNA and protein levels maximally accumulating 48-96 h post-IR. The low levels of IR (>0.02 Gy) that induce sCLU and the dramatic accumulation of sCLU protein following taxol, TPA or thapsigargin (a sarcoplasmic reticulum Ca2+ -ATPase

58 (SERCA) pump inhibitor that causes dramatic alterations in intracellular calcium homeostasis) exposures, suggest that DNA damage may not be required for CLU gene expression in MCF-7 breast cancer cells. Expression of the human papillomavirus (HPV)

E6 protein, as well as isogenically matched cell lines that differ only in their p53 status, were used to demonstrate a role for p53 in the transcriptional repression of sCLU in unirradiated as well as IR-exposed cells. Loss of functional p53 results in elevated basal levels of sCLU in some cells, and augmented IR-induced gene expression in all cells examined. These data strongly suggest that sCLU mRNA production and protein synthesis are repressed by the tumor suppressor protein, p53.

EXPERIMENTAL PROCEDURES

Chemical Reagents

Camptothecin, etoposide, colcemid, nocodazole, taxol, mimosine, TPA, and thapsigargin were obtained from the Sigma Chemical Co (St. Louis, MO) and dissolved in either PBS or DMSO. Topotecan was generously provided by Glaxo SmithKline (Research Triangle

Park, NC). ß-Lapachone was prepared for us by Dr. William G. Bornmann (Synthetic

Preparatory Core Facility, Memorial Sloan Kettering, NY, NY). Taxotere was generously provided to us by Aventis Pharmaceuticals (Bridgewater, NJ).

Cell Culture

MCF-7:WS8 human breast cancer cells (MCF-7) were obtained from Dr. V. Craig Jordan

(Northwestern University; Evanston, IL). MCF-7 cells were transduced by retroviral transfer with a CMV-driven papillomavirus E6 vector by Dr. Jordan's lab, and

59 subsequently subcloned by our lab into cell lines with varying E6 expression. The E6-D

MCF-7 cell line showed no p53 expression, even after IR exposure. Human colorectal carcinoma HCT116 parental, p53-/-, and p21-/- cell lines were developed (321) and generously provided by Dr. Bert Vogelstein (Johns Hopkins University; Baltimore, MD).

These cell lines were confirmed by our laboratory to be null for p53 and p21, respectively, by western blot analyses. MCF-7, ZR-75-1, T47-D, BT474, MDA-MB-231 and MDA-MB-468 cell lines were grown in RPMI 1640 cell culture media supplemented

with 5% fetal bovine serum (FBS) at 37°C in a humidified incubator with a 5% CO2-95% air atmosphere as described (322). MCF-7:E6 cells were maintained in 0.4 mg/ml geneticin (Life Technologies; Carlsbad, CA). RKO:neo and E6-expressing RKO cell lines were obtained from the American Type Culture Collection, maintained in G418, and experiments performed in the absence of antibiotics. HCT116 and RKO cell lines were grown in DMEM supplemented with 10% FBS at 37°C in a humidified incubator with a

5 10 % CO2-90% air atmosphere. All experiments were initiated by seeding 5 x 10 log- phase cells per 10-cm2 tissue culture dish in the appropriate medium in the absence of any antibiotics (e.g., geneticin). All cell lines were free from mycoplasma contamination.

IR and Chemical Treatments

Cells were irradiated as described (11). Briefly, cells were irradiated with 137Cs gamma rays at a dose rate of 0.87-0.92 Gy/min, using a Shepard Mark Irradiator. Untreated cells were mock-irradiated as described (11). MCF-7 cells were treated with topotecan, camptothecin, mimosine, colcemid, nocodazole, TPA, thapsigargin, taxotere, taxol or etoposide using drug exposures at the indicated doses as described in Table 2. Cells were

60 treated with ultraviolet radiation as described (91). ß-Lapachone and hypoxic exposures

(323) of log-phase MCF-7 cells were performed as indicated in Table 2.

Northern Blot Analyses

Total RNA was extracted from control or irradiated MCF-7 or HCT116 cells as indicated using Trizol (Life Technologies; Carlsbad, CA) as per the manufacturer’s instructions.

Total RNA (10 - 20 µg) was separated on a denaturing formaldehyde gel, transferred to a

Hybond membrane (Amersham Pharmacia; Sunnyvale, CA), and probed with 32P-labeled full-length CLU or 36B4 cDNAs as described (91); 36B4 levels are not affected by cell stress, or cell cycle status (91). Corresponding transcript signals were quantified using

ImageQuant software version 4.1 (Molecular Dynamics; Sunnyvale, CA) on a Molecular

Dynamics phosphoimager. CLU mRNA levels were normalized to untreated control levels, and to 36B4 mRNA levels for X-fold induction calculations as described (91).

Luciferase Assays

All luciferase assays were performed using the Luciferase Assay System (Promega;

Madison, WI). MCF-7 cells were stably transfected with a 1403 bp fragment of the human CLU promoter in a luciferase reporter plasmid using a standard liposome transfection protocol (Effectene, Qiagen, Valencia, CA). The plasmid was a generous gift from Dr. Martin Tenniswood (University of Notre Dame; Notre Dame, IN). These cells (MCF-7:1403 cells) were seeded in 6-well plates at approximately 50% confluency.

Cells were irradiated at the indicated dose(s) and harvested at various times in 1X

61 reporter lysis buffer (Promega; Madison, WI). Each dose/time point was completed in triplicate and a Student’s T-Test was performed to determine statistical significance.

Western Blot Analyses

Whole cell extracts from control or irradiated cells were extracted in RIPA buffer (0.1%

SDS, 0.5% deoxycholate, 1% NP-40, 150 mM NaCl, 50 mM Tris pH 8.0) and separated on a 10% gel by SDS-PAGE western blot analyses as described (35). Proteins were transferred to Immobilon-P (Millipore; Bedford, PA) and probed with the B-5 human sCLU monoclonal antibody, the DO-1 human p53 monoclonal antibody, and the human

C-19 Ku70 polyclonal antibody. All antibodies were obtained from Santa Cruz and used as per manufacturer’s instructions. Ku70 was used as a control for equal loading of protein, since its levels remain unaltered after IR or cell cycle status under the time-frame of our experiments. Western blots shown are representative of experiments performed at least three times. For sCLU protein analyses in cells before and after IR, we routinely use the 60 kDa form, since all cell lines produce this 60 kDa protein, which is a precursor to the mature glycosylated α- and β- ~40 kDa polypeptides of sCLU.

Cell Cycle Analyses

HCT116 parental, p53-/- or p21-/- cells were synchronized by allowing them to grow to

100% confluence on 10-cm2 tissue culture dishes as described (11, 324). Cells were then

maintained for 48 h in serum-free medium to maximize G0-G1 arrest. Cells were released from the dual confluence and serum-free cell cycle arrest by trypsinization (using 0.05% trypsyn with 0.53 mM EDTA) and replated at 1:8-1:10 dilution in DMEM containing

62 10% FBS under conditions described above. For IR treatments, cells were exposed to 10

Gy, 10 h after release from the cell cycle. Concurrent flow cytometric and western blot analyses were performed as indicated. At various times after mock- or IR-exposures, cells were dissociated by scraping into 1X PBS, collected by centrifugation (500 x g), fixed in

90% ethanol, and stored at –20 0C until analyzed. Cells were then stained with 33 mg/ml propidium iodide (PI) (Sigma; St. Louis, MO), 1.0 mg/ml RNase A (Sigma), and 0.2%

NP-40 (Calbiochem; La Jolla, CA) at 4 0C overnight. Stained nuclei were then analyzed for DNA content by PI fluorescence using a Coulter Epics XL (Beckman Coulter

Electronics; Miami, Fl) flow cytometer. Data were analyzed using ModFit LT, version

2.0 software (Verify Software House; Topsham, ME). Western blot analyses were completed simultaneously with flow cytometry, and results shown represent experiments performed at least three times.

RESULTS sCLU is transcriptionally upregulated after IR.

Our laboratory previously showed that CLU was an x-ray-induced protein (xip8) (91), however, regulation of the CLU gene was not elucidated. To further characterize induction of sCLU after IR exposure, northern blots were used to determine if sCLU protein accumulation in log-phase MCF-7 human breast cancer cells after IR exposure was due to increased transcription, or a result of protein stabilization (Figures 2.1A,B).

Log-phase MCF-7 cells were mock-irradiated or exposed to 10 Gy and harvested at various times post-IR to analyze the temporal kinetics of CLU gene expression. Maximal induction of CLU mRNA (7- to 10-fold) over untreated cells occurred 72 to 96 h after 10

63 Gy (Figure 2.1A). Induction of the CLU transcript in MCF-7 cells after IR was confirmed using RNase protection assays (data not shown). IR dose-response experiments in log-phase MCF-7 cells were performed, and CLU mRNA accumulations

72 h after exposure were examined (Figure 2.1B), since maximal protein and mRNAs were noted at this time post-IR (Fig. 2.1A). As previously reported for sCLU protein induction at 72 h post-IR (35). CLU mRNA was induced 2-fold after as little as 2 cGy, with maximal induction of 22-fold in MCF-7 cells after 5 Gy (Figure 2.1B). Steady state

CLU mRNA accumulation corresponded well with previously described sCLU protein accumulation in MCF-7 cells after IR in terms of temporal and dose-response kinetics

(35).

To determine if IR-induced CLU transcriptional increases were due to de novo mRNA synthesis, or to decreases in mRNA degradation (i.e., via post-transcriptional modifications), we examined CLU promoter activity in time-course and dose-response studies after IR. For these experiments, we generated an MCF-7 cell line containing a stably integrated copy of a plasmid containing a 1403 bp fragment of the human CLU promoter directing expression of a downstream luciferase reporter as described in

'Experimental Procedures'. Transient transfections with the CLU reporter plasmid were problematic, since all transfection methods examined to date affected the regulation of the CLU promoter-luciferase construct in MCF-7 cells, as well as endogenous sCLU gene/protein expression (data not shown); induction of sCLU may be triggered by cell membrane insult (325). Dose-response (Figure 2.1C) and time-course (Figure 2.1D) assays of exogenous CLU promoter activation in MCF-7:1403 cells were performed to show that this clone behaved similarly to the endogenous CLU gene before and after IR

64

C A 1400 UT 10 Gy * 1200 *p<0.03 Time (h): 96 4 24 48 72 96 1000 800 1.8 kb CLU 600 * 400 * 0.7 kb 36B4 200 0 Fold 1 Induction: 1 1 2 8 7 8 4 24 48 72 96 4 24 48 72 96 Relative Luciferase Units UT 10 Gy D B 7.E+03 * X-irradiation (Gy) 6.E+03 *p<0.02 UT 0.02 0.05 0.1 0.25 0.5 1.0 2.5 5.0 10 5.E+03

4.E+03 * 1.8 kb CLU 3.E+03 * 2.E+03 0.7 kb 36B4 * 1.E+03 Fold 0.E+00 Induction: 1 2 2 4 5 7 11 13 22 7 1 0 0.5 2 5 10 Relative Luciferase Units X-irradiation (Gy)

65 Figure 2.1. sCLU is transcriptionally upregulated after IR exposure in MCF-7 human breast cancer cells. CLU mRNA levels were monitored in asynchronous

MCF-7 cells after 10 Gy by northern blot analyses and luciferase assays. In A, log- phase growing MCF-7 cells were irradiated with 10 Gy and 10 µg of total RNA was analyzed by northern blot analyses as described in Experimental Procedures. In B,

MCF-7 cells were irradiated with various doses of IR and total RNA was harvested

72 h after exposure. Total RNA (10 µg) was used for northern blot analyses. In C, time-course of sCLU induction after 10 Gy exposure was analyzed by luciferase assays in MCF-7 cells stably transfected with 1403 base pairs of the CLU promoter

(i.e., MCF-7 1403 cells) using the Luciferase Assay System (Promega). In D, an IR dose-response was performed on the MCF-7 1403 cells 72 h after IR exposure. Each dose/time point was performed in triplicate and a Student’s T-Test was performed to determine statistical significance. *p values denote IR compared to untreated

(UT) controls.

66 exposure. The CLU promoter was activated in a time- and dose-dependent manner similar to that previously shown for sCLU protein and mRNA (Figures 2.1A and B).

CLU promoter activity was stimulated by as low as 0.5 Gy.

sCLU is a stress protein induced by a variety of cytotoxic agents.

Table 2.1 lists various cytotoxic agents that induce sCLU protein expression in MCF-7 cells. These agents included ultraviolet radiation (UV), topoisomerase I and IIα poisons, microtubule stabilizers/destabilizers, as well as other agents that do not cause direct damage to DNA (e.g., TPA, thapsigargin). Treatment of MCF-7 cells with hypoxic conditions or various doses of β-lapachone (2-10 µM, 4h), a novel apoptotic drug that quickly depletes cellular NAD(P)H and ATP in NQO1-expressing MCF-7 cells (326), did not induce sCLU protein expression. These data suggest that damage to DNA may be sufficient, but is not required for sCLU induction. Alterations in calcium homeostasis

(indicated by thapsigargin induction of sCLU, Table 2.1) or ER stress responses may play a common role in triggering CLU gene induction.

Correlation of sCLU expression and loss of functional p53.

Various human breast, colon and prostate cancer cell lines with known mutations in p53 were examined for basal and IR-inducible sCLU levels as monitored by western blot analyses and described in 'Experimental Procedures'. With one exception, cells expressing mutant p53 exhibited increased basal levels of sCLU (Table 2.2). Mutant p53-expressing MDA-MB-231 cells appear to lack basal or IR-inducible CLU protein expression, and have no detectable CLU mRNA levels by Northern blot analyses (data

67 not shown). In contrast, cells expressing wild-type p53 expressed low or no detectable basal levels of sCLU (Table 2.2). With the exception of MCF-7 cells, we also noted that cells expressing wild-type p53 did not greatly induce sCLU expression after various doses of IR to the same extent as null or mutant p53-expressing cells.

The data in Table 2.2 indicate an inverse correlation between sCLU expression and expression of wild-type p53. To further elucidate the effect of p53 on sCLU expression, we compared vector alone-transfected parental MCF-7 cells to isogenically matched cells stably transfected with the HPV-16 E6 protein, as described in

‘Experimental Procedures’. The E6 protein binds to p53 and targets it for rapid degradation through the proteasome pathway, leaving these cells deficient (i.e., null) for p53 expression (327). Protein and RNA from log-phase MCF-7:parental and MCF-7:E6 cells were harvested at various times after exposure to 10 Gy. Consistent with the mRNA changes shown in Fig. 2.1, sCLU protein was induced in parental MCF-7 cells starting at

24 h, and levels peaked at 72 h post-10 Gy (Figure 2.2A). The 60 kDa band in Figure

2.2A is a precursor form of sCLU (psCLU) expressed in the ER and is cleaved at an α/β cleavage site resulting in two 40 kDa peptides that heterodimerize through five disulfide bonds to form mature 80 kDa sCLU protein. Western blots performed under denaturing conditions result in the appearance of a 40 kDa smeared band consisting of glycosylated

α- and β-peptides of sCLU. sCLU basal levels were higher in mock-irradiated MCF-

7:E6 cells compared to parental MCF-7 cells. As expected, p53 protein levels were not detected in MCF-7:E6 cells at various times before or after IR. Furthermore, induction of sCLU protein was enhanced in MCF-7:E6 cells after IR compared to levels in parental

68 Table2.1

Induction of sCLU protein expression in MCF-7 cells1 by various cytotoxic agents.

Agent Dose Range for Induction2 Level of Induction3 DNA Damaging Agents Ionizing Radiation (IR) 0.02 - 10 Gy +++ Ultraviolet Radiation (UV) 12 J/m2 +++ Topotecan 50 nM +++ Camptothecin 100 nM +++ Etoposide (VP-16) 15 µM +++ Non-DNA Damaging Agents Photodynamic Therapy (PDT)4 200 nM PC-4/200 mJ/cm2 ++ Colcemid 70 ng/ml ++ Nocodazole 150 ng/ml +++ Taxol 1 - 50 nM +++ Taxotere 1 - 10 nM +++ Mimosine 0.5 mM ++ TPA 100 nM + Thapsigargin 10 - 500 nM ++ Non-Inducing Agents β-Lapachone 2 - 10 mM NA 5 Hypoxia (<0.1% O2) NA

1Log phase MCF-7 cells were seeded at approximately 5 X 105 cells per 10 cm plate.

2Topotecan, camptothecin, TPA and thapsigargin were continuous treatments.

Cells were treated with colcemid, nocodazole and mimosine for 24 hours, washed with PBS and replated into fresh media. Cells were treated with taxol or taxotere for 4 hours, washed with PBS and replated into fresh media. Cells were treated with VP-16 for 1 hour. Protein was harvested at least 48 hours after drug addition/irradiation.

3Level of induction of sCLU protein as compared to sCLU protein derived from 10

Gy irradiated MCF-7 cells.

69 4Photosensitizing drug used was Phthalocyanine 4 (PC-4). Induction of sCLU protein was only seen after addition of drug and light exposure. No induction was observed with light alone or PC-4 alone.

5Hypoxia was induced as previously described (323).

70 MCF-7 cells after 10 Gy (Figure 2.2A). Since MCF-7:E6 cells have a higher basal level of sCLU, IR-induction of sCLU in these cells was more difficult to quantify, and

Northern

HPV-16 E6-expressing MCF-7 cells have high basal levels of sCLU. blot analyses indicate this fact (see below). Finally, we noted a similar dramatic increase of sCLU in the medium of IR-treated MCF-7 cells, and a significantly higher basal level of sCLU in the medium of MCF-7:E6 compared to MCF-7:neo vector alone parental cells (Klokov et al., unpublished data).

Northern blot analyses confirmed that basal CLU mRNA levels were 3-fold higher in mock-treated MCF-7:E6 cells compared to MCF-7:parental cells (Figure 2.2B).

Furthermore, CLU mRNA levels were only modestly induced in MCF-7:E6 cells (~3-fold from 4h to 96 h post-IR), compared to IR-treated vector alone MCF-7:parental cells, in which a 10-fold increase in CLU mRNA level was noted (Figure 2.2B).

RKO:neo and RKO cells stably expressing the E6 protein (RKO:E6) were mock- irradiated or treated with 5 Gy. Protein was harvested 72 h after exposure and analyzed by western blotting (Figure 2.2C). As with many other cells examined, RKO cells showed low levels of the mature 40 kDa glycosylated form of sCLU, presumably because this protein is secreted from the cell. Wild-type p53-expressing RKO:neo cells expressed low basal levels of sCLU protein, with a measurable IR-inducible expression of the 60

71 Table 2.2

Effect of p53 status on sCLU basal and IR-inducible expression.

IR Cell Line p53 Status Basal1 Inducibility2 RNA3 Breast cancer cell lines MCF-7:parental wild-type (wt) low yes + MCF-7:E6D wt (no expression) high yes + ZR-75-1 wt low no + T47-D mutant (194) high no + BT474 mutant (275) high no + MDA-MB-231 mutant (280) ND ND ND MDA-MB-468 mutant (273) high no + Colon cancer cell lines HCT116:parental wt low minimal + HCT116:p21 -/- wt low minimal + HCT116:p53 -/- null low yes + RKO:neo wt low minimal NP RKO:E6 wt (no expression) high yes NP Prostate cancer cell lines LNCaP wt low minimal NP DU-145 mutant (275) high minimal NP

1Basal levels were determined as compared to log-phase growing untreated MCF-7 parental cells.

2Log-phase growing cells were treated with 10 Gy IR and protein was harvested 48 hours after exposure. MCF-7 parental cells were used as the standard for “high” IR inducibility.

3RNA status was determined by RT-PCR using primers designed to full length CLU

DNA.

ND, not detected; NP, not performed

72 kDa sCLU protein form after IR. In contrast, RKO:E6 cells expressed high basal levels of the 60 kDa sCLU protein, similar to that seen in MCF-7:E6 cells. As in MCF-7 cells, sCLU induction was more dramatic in RKO:E6 cells compared to RKO:neo cells, suggesting that loss of p53 function relieves IR-induction responses of sCLU.

Somatic deletion of p53 in HCT116 colon cancer cells results in greater IR-inducible sCLU levels.

Since E6 expression may have additional unknown ‘gain of function’ properties, and to confirm the ability of p53 to repress sCLU expression after IR, we used isogenically matched human HCT116 colon cancer cell lines that differed only in their p53 or p21 status. As with RKO cells, we were not able to observe intracellular mature 40 kDa sCLU levels in HCT116 cells and IR-induced sCLU induction responses were monitored via the ~60 kDa sCLU precursor protein (psCLU). Protein and RNA from log-phase

HCT116:parental and HCT116:p53-/- cells were harvested at various times after exposure to 10 Gy. Western blot analyses showed that IR-treated HCT116:parental cells stabilized and accumulated p53 (i.e., expressed wild-type p53), but increases in steady state levels of sCLU were minimal to non-detectable (Figure 2.3A). In contrast, HCT116:p53-/- cells dramatically induced sCLU after 10 Gy. As found with RKO:E6 and MCF-7:E6 cells,

HCT116:p53-/- cells expressed higher basal levels of sCLU compared to the low levels

73 MCF-7:parental MCF-7:E6D A UT 10 Gy UT 10 Gy Time (h): 4 96 4 24 48 72 96 4 96 4 24 48 72 96 psCLU

sCLU

p53

Ku70

B UT 10 Gy UT 10 Gy Time (h): 4 4 24 48 72 96 4 4 24 48 72 96

1.8 kb CLU

0.7 kb 36B4

Fold Induction: 1 1 3 8 10 9 3 3 3 5 8 9

C RKO:neo RKO:E6 UT 5 Gy UT 5 Gy psCLU

p53

Ku70

Figure 2.2. sCLU basal levels are elevated in MCF-7 and RKO cells that over- express the HPV E6 protein. MCF-7:parental and MCF-7:E6 cells were exposed to

10 Gy and protein was harvested at various time points. In A, protein (100 µg) was loaded for each sample and separated by standard 10 % SDS-PAGE. Blots were probed for sCLU, p53 and Ku70 using western blot analyses as described in

Experimental Procedures. Ku70 was used as a loading standard as described. In B, total RNA (10 µg) was analyzed using standard northern blot techniques as

74 described in Fig. 2.1. Shown are representative blots from experiments performed at least three times. In C, RKO:neo and RKO:E6 cells were exposed to 5 Gy and protein was harvested at 72 h. Protein (100 µg) was loaded for each sample and separated by standard 10 % SDS-PAGE. Blots were probed for sCLU, p53 and

Ku70 using western blot analyses as described in Experimental Procedures.

PsCLU, presecretory 60 kDa clusterin form; sCLU, mature secretory 80 kDa clusterin form seen as 40 kDa glycosylated α - and β- polypeptides. Ku70 is a nonhomologous end joining protein that is not altered by cell cycle or cell stress changes.

75 noted in HCT116:parental cells. IR dose-response analyses of sCLU responses in

HCT116:p53-/- cells demonstrated induction of sCLU protein at doses as low as 1 Gy

(Figure 2.3B), a dose of IR that caused minimal loss of clonogenic survival (97).

Northern blot analyses confirmed induction (6- to 7-fold) of steady state CLU mRNA in

HCT116:p53-/- cells (Figure 2.3C), whereas p53+/+ HCT116:parental cells showed little or no induction of CLU mRNA at various times (up to 96 h) after 10 Gy.

To demonstrate that induction of sCLU in p53 null cells was specific for the absence of p53 and not a gene downstream from p53, we utilized HCT116 cells that were somatically knocked out for the p21 gene (i.e., HCT116:p21-/- cells) (321). As in p53+/+

HCT116:parental cells, sCLU protein and mRNA levels were only minimally induced in

HCT116:p21-/- cells after 10 Gy compared to mock-treated cells, as determined by western and northern blot analyses (Figure 2.4A, B).

76 A HCT116:parental HCT116:p53-/- UT 10 Gy UT 10 Gy Time (h): 4 96 4 24 48 72 96 4 96 4 24 48 72 96

sCLU

p53

Ku70

B X-irradiation (Gy) UT 1.0 2.0 5.0 10

sCLU

Ku70

C UT 10 Gy UT 10 Gy 96 4 24 48 72 96 96 4 24 48 72 96

sCLU 1.8 kb

0.7 kb 36B4

Fold Induction: 1 <1 <1 <1 <1 1.5 1 1 1 3 6 7

Figure 2.3. sCLU is induced in HCT116:p53-/- cells, but not in p53+/+

HCT116:parental cells. Asynchronous HCT116:parental and HCT116:p53-/- cells were exposed to 10 Gy and protein harvested at various times. In A, western blot analyses were performed as in Figure 2.2, and probed for sCLU, p53 and Ku70 as described in Experimental Procedures. In B, HCT116:p53-/- cells were treated with various doses of IR and protein was harvested 72 h later. Western blot analyses were performed as in Figure 2.2. In C, total RNA (10 µg) was analyzed using

77 northern blot techniques as described in Figure 2.1 and Experimental Procedures.

Shown are representative blots from experiments performed at least three times.

78 sCLU is not cell cycle regulated.

An alternative explanation for sCLU induction and subsequent repression by p53 could be that the CLU gene is cell cycle regulated, and that wild-type p53-expressing cells suppress sCLU expression by arresting cells in a particular phase of the cell cycle.

Recent reports suggested that sCLU may be expressed exclusively in quiescent normal cells (328). To address this issue in cancer cells, HCT116:parental, HCT116:p53-/- and

-/- HCT116:p21 cells were arrested in the G0/G1 phase of the cell cycle by dual serum- starvation and confluence-arrest conditions, released by replating and irradiated 10 h later as described in ‘Experimental Procedures’ (329). The cell cycle profiles of untreated and irradiated isogenic HCT116 cells were then monitored (Figure 2.5). Untreated

HCT116:parental cells subsequently entered S-phase 14-16 h after release from low

serum and confluence arrest, with concomitant decreases in G0/G1 cells.

As previously noted with this synchronization technique (330), one synchronous cell division was achieved, and mock-irradiated isogenic HCT116 cells returned to a log- phase cell cycle distribution after 58 h post-release. There were significant differences in synchronized mock-irradiated HCT116:parental, HCT116:p53-/- and HCT116:p21-/- cells, particularly in the time of entry into S-phase, with both p53- and p21-deficient HCT116 cells entering S-phase sooner than wild-type p53 HCT116:parental cells (compare the

cell cycle distributions in Figures 2.5A, C and E). Since p53 exerts its G1 cell cycle checkpoint responses through, in part, induction of p21, mock-irradiated and IR-treated

HCT116:p53-/- and HCT116:p21-/- cell cycle distributions were very similar (Figures

2.5C, E).

79 A HCT116:p21-/- UT 10 Gy Time (h): 96 4 24 48 72 96

sCLU

p53

Ku70 B UT 10 Gy Time (h): 4 96 4 24 48 72 96

sCLU 1.8 kb

36B4 0.7 kb

Fold Induction: 1 <1 <1 <1 <1 <1 <1

Figure 2.4. sCLU is not induced in HCT116:p21-/- cells. HCT116:p21-/- cells were exposed to 10 Gy and protein harvested at various times. In A, western blot analyses were performed as in Figure 2.2. Blots were probed for sCLU, p53 and

Ku70 by western blot analyses as described in Experimental Procedures. In B, total

RNA (20 µg) was analyzed using standard northern blot techniques as described in

Figure 2.1. Shown are representative blots from experiments performed at least three times.

80 As expected, p53+/+ HCT116:parental cells treated with 10 Gy at 10 h post release resulted in a significant delay in the progression of synchronized cells into S-phase (a

function of the IR-induced G1 cell cycle checkpoint response, Figures 2.5A, B) as described (330). For example, at 36 h IR-treated HCT116:parental cells demonstrated

>45% G2 cells compared to less than 18% in mock-irradiated cells (compare Figures

2.5A, B). In contrast, HCT116 cells with somatic deletions of p53 (Figure 2.5C, D) or p21 (Figures 2.5E, F), entered S-phase earlier, with accumulation of S-phase cell

populations occurring at 12-16 h, accompanied by concomitant decreases in G0/G1 cells.

At 18 h after release (8 h after 10 Gy IR exposure), only 14% and 28% of p53-/- and p21-/-

cells, respectively, remained in G1, while 80% and 66% of cells, respectively, proceeded

+/+ into S phase. As expected, IR-treated HCT116:p53 cells arrested in G1 and exhibited

-/- -/- delayed S or G2/M phase entry compared to IR-treated HCT116:p53 or HCT116:p21 cells. HCT116:p53-/- and HCT116:p21-/- cells responded similarly to IR treatment.

Although responses to 10 Gy are shown, near identical responses to 2-5 Gy were also observed in other studies (330).

Western blot analyses of non-irradiated synchronized HCT116:parental,

HCT116:p53-/- and HCT116:p21-/- cell populations indicated that the levels of sCLU did not change relative to basal levels throughout the cell cycle (Figure 2.5). Interestingly, sCLU was induced only in IR-exposed synchronized HCT116:p53-/- cells with similar induction kinetics (maximal accumulation observed between 24-72 h) as noted in IR- treated asynchronous log-phase HCT116:p53-/- cells (Figure 2.3). In contrast, only minimal sCLU induction responses were noted in synchronized IR-treated

HCT116:parental or HCT116:p21-/- cells, even though irradiated HCT116:p21-/- cells

81 exhibited nearly identical cell cycle distribution changes as IR-exposed HCT116:p53-/- cells. These data strongly suggested that: (a) sCLU induction was genetically programmed after IR stress resulting in a 48 – 72 h delay before sCLU accumulation was noted after IR exposure; and (b) sCLU was transcriptionally repressed by functional p53 independent of the cell cycle. Loss of functional p53 appears to relieve negative regulation on the IR-induction responses of CLU gene expression in a variety of cell types.

DISCUSSION

Our laboratory previously demonstrated that sCLU was an x-ray induced protein (i.e., xip8) (35). In this study, we further investigated the induction of sCLU by IR. We showed that sCLU was induced by doses of IR as low as 2 cGy (Figures 2.1A & B).

This low-dose IR induction is seen at both the transcript and protein levels, with promoter activation noted after 0.5 Gy in MCF-7 cells containing a stably integrated 1403 bp human CLU promoter directing expression of firefly luciferase. Induction of the CLU promoter was noted only after 0.5 Gy as analyzed by luciferase assays using a luminometer (Figures 2.1C & D), however, induction of this promoter after IR doses <50 cGy has been noted when analyzed by bioluminescent imaging (Klokov et al., unpublished data). We also showed that DNA damage appears to be sufficient, but not required for sCLU induction (Table 2.1).

We demonstrated that the basal level and IR induction of sCLU after IR exposure was repressed by p53. MCF-7 and RKO cells stably expressing the HPV E6 protein

(both exhibiting loss of functional p53) have high basal levels of sCLU compared to

82 parental cells that express functional p53 (Figure 2.2). Additionally, HCT116:parental and RKO:neo cells that express wild-type p53 minimally induce sCLU after IR, whereas

HCT116:p53-/- and RKO:E6 greatly induced sCLU at the protein and transcript level after

IR (Figure 2.3). Finally, the effect of p53 on IR-inducible sCLU expression was not dependent on the cell cycle, but was delayed in its induction, requiring at least 48 h post-

IR in all cells examined (Figure 2.5). Since the relationship between p53 status and sCLU expression and p53 repression of IR-induced sCLU expression was observed in cells of different origins, p53 repression of this gene appears to be a general phenotype and not unique to specific cell lines. A limited screen of cancer cell lines indicated an inverse regulatory relationship between p53 status and sCLU expression (Table 2.2). To directly explore the role of p53 in basal and IR-inducible levels of CLU gene expression, we used three model cell line systems from breast and colon cancer origins to investigate the role of p53 in the transcriptional regulation of sCLU. All three cell lines confirmed that p53 exerted negative regulation on sCLU expression.

The effect of IR exposure on sCLU expression in MCF-7 cells was different from that found in HCT116 parental cells, even though both cell lines express wild-type p53. HCT116:parental cells did not induce sCLU after IR exposure. In fact,

MCF-7 cells are the only wild-type p53-expressing cell line examined to date that strongly induced sCLU after IR. It may be that MCF-7 cells over-express the IR- activated transcription factors required for induction of CLU gene expression, while

HCT116 cells maintain lower levels, which are in turn efficiently suppressed by wild- type p53 even after IR exposures. It appears that these as yet unknown transcription

83 A B HCT116:parental 100 parental 100 IR UT 80 80 60 60 40 40 20 20 0 0 0 10 20 30 40 50 60 0 10 20 30 40 50 60 Percent Cell Number Time (h) after 5 Gy MCF-7 0 10release 18 (h) 24 36 58 5 Gy: - + - - - + - + - + - + CLU

p53

Ku70

C HCT116:p53-/- D 100 -/- 100 IR UT p53 80 80 60 60 40 40 20 20 0 0 0 10 20 30 40 50 60 0 10 20 30 40 50 60 Percent Cell Number Time (h) after 5 Gy MCF-7 0 release 10 18 (h) 24 36 58 5 Gy: - + - - - + - + - + - + CLU

p53

Ku70

HCT116:p21-/- E p21-/- F 100 IR 100 UT 80 80 60 60 40 40 20 20 0 0 0 10 20 30 40 50 60 0 10 20 30 40 50 60 Percent Cell Number Time (h) after 5 Gy MCF-7 0 release 10 18 (h)24 36 58 5 Gy: - + - - - + - + - + - +

CLU

p53

Ku70

84 Figure 2.5. sCLU is not cell cycle regulated. HCT116:parental, p53-/- and p21-/- cells were synchronized by serum starvation and confluence-arrest. Synchronized cells were released by low density seeding in 10% FCS-DMEM medium and cells were allowed to proceed through the cell cycle for 10 h and then mock-irradiated

(Fig. 2.5A, C, E), or exposed to 10 Gy (Fig. 2.5B, D, F). Cells were allowed to

progress through G1 (), S () and G2/M () phases of the cell cycle. Protein was harvested for flow cytometric or western blot analyses at the indicated times as described previously in Experimental Procedures. Western blots were probed for

CLU, p53 or Ku70 expression as described in Experimental Procedures. Shown are data for HCT116:p53+/+ parental (A, B), HCT116:p53-/- (C, D) and HCT116:p21-/- (E,

F) cells. Western blots and cell cycle analyses are representative of experiments performed at least three times.

85 factors may be constitutively expressed in MCF-7 cells, since E6 expression greatly enhanced sCLU expression in MCF-7 cells without IR exposure, whereas loss of functional p53 in HCT116 cells did not cause an appreciable increase in basal levels of sCLU protein expression. The factors needed for sCLU induction have not been elucidated. Analyses of the transcription factors and DNA elements within the CLU promoter that regulate the IR inducibility of this gene are currently being performed in our laboratory.

The signaling pathway(s) that regulate sCLU induction and expression after IR exposure is(are) unknown. Our laboratory identified CLU as a Ku70 binding protein using yeast-two-hybrid analyses (71). Through our screen of cytotoxic agents, we noted that DNA damage was not required for sCLU induction. This was best demonstrated by the induction of sCLU after thapsigargin (TG) exposures, and at doses of TG (2 nM, 1 h) that are not lethal to exposed MCF-7 cells (Table 2.1). TG is an inhibitor of the SERCA pump in the ER. Treatment of MCF-7 cells with TG resulted in a transient release of intracellular calcium (331) and an induction of sCLU mRNA and protein, suggesting that calcium changes may be an upstream signaling event mediating sCLU induction. It is possible that calcium, as a signaling molecule, may be a triggering event common to all the agents in Table 1 that elicit sCLU induction responses. The exact signal transduction processes that result in CLU gene expression after DNA damaging agents compared to non-DNA damaging agents is being elucidated in our laboratory.

The mechanism of sCLU repression by p53 also remains to be elucidated. There are several proposed models of p53 transcriptional repression. In the first model, p53 binds to its putative DNA binding sequence and sterically inhibits the binding of

86 transcription factors required for induction. This model was proposed to account for repression of Bcl-2 (332), α-fetoprotein (333) and HBV (334) genes by p53. In the second model, p53 binds and sequesters transcription factors required for upregulation.

For example, p53 can directly bind several transcription factors including Sp-1 (335,

336), AP-1 (337), NF-Y (338, 339), Brn-3a (332) and C/EBPβ (340), that may be responsible for upregulated CLU promoter activity after IR. Additionally, it was shown that p53 can bind the TATA binding protein (TBP) in vitro and inhibit transcription by disrupting formation of the TFIID complex (341). Alternatively, Johnson et al. have proposed a novel putative DNA binding sequence for p53 that is strictly involved in transcriptional repression (342).

Collectively, our data strongly suggest that the CLU gene is transcriptionally repressed by p53, although the mechanism of this repression still remains to be elucidated. The cell models used in this study will allow us to further investigate the mechanism(s) of p53 repression of sCLU, as well as the signaling pathways required for sCLU induction after IR exposure. Understanding the cellular responses to ionizing radiation exposure, in normal and tumor tissue, is vital for improving the efficacy of radio-therapy in the clinic.

The data presented in this paper provide a first examination of how a cell may regulate the clusterin molecular switch, turning on the cytoprotective sCLU protein at low doses of IR (0.02 - 0.1 Gy), while at the same time allowing p53 responses after high doses of IR (≥ 1.0 Gy) to shut down this cytoprotective protein to allow for cell cycle checkpoint responses and for cell death in severely damaged cells. For example, we are exploring the possibility that functional p53 is responsible for mediating CLU alternative

87 splicing that produces nCLU protein expression (95). We previously demonstrated that nCLU, and not sCLU, could associate with Ku70 and cause apoptotic cell death responses. In this way, p53 would down-regulate the cytoprotective sCLU protein, while simultaneously stimulating the synthesis and possibly activation of nCLU. Expression, and nuclear translocation, of nCLU after >1 Gy of IR would then result in a cascade of events leading to cell death and apoptosis. Understanding the regulatory events affecting the relative levels of different forms of the CLU protein after IR should allow elucidation of ways to modulate death responses in tumor cells, while possibly sparing the survival of normal cells.

88 Chapter 3: Induction of Clusterin, a Pro-Survival Factor, Requires Delayed

Activation of IGF-1R/MAPK Signal Cascade after IR

Submitted to Molecular Cell, March 2004

Abstract

Radiation therapy is a common treatment for many tumor types, yet the responsiveness to this treatment is variable. We show that secretory clusterin (sCLU), an ionizing radiation

(IR) – induced gene provides cytoprotection after IR. Induction of sCLU is a delayed event taking 72 hrs after radiation to reach maximal levels. This delayed induction is dependent upon the up-regulation and activation of insulin-like growth factor-1 receptor

(IGF-1R) and the downstream Src-Raf-Mek-Erk signaling cascade. This pathway culminates in the activation of early growth response-1 (Egr-1) that is required for sCLU induction. Thus, our data provides a mechanism by which cancer cells may provide cytoprotection for neighboring cells through potential bystander effects, and the development of radio-resistance after clinically relevant doses of IR through the IGF-1R-

Src-Mek-Erk-Egr-1 cascade.

89 Introduction

Ionizing radiation (IR) is a common therapy for many types of , and elucidation of refractory responses to clinically relevant doses of IR are under intense investigation.

Understanding these cellular processes, especially the contribution of signal transduction pathways that precede gene expression after IR will lead to interventions that bring about a more favorable clinical outcome. Although it was once thought that the only important cellular responses to IR originated from DNA damage, it is clear that IR creates many different lesions in macromolecular targets within the cell that set in motion cascades of responses in both irradiated, as well as in neighboring non-irradiated cells. For example, the epidermal growth factor receptor (EGFR) (343, 344) , insulin-like growth factor-1 receptor (IGF-1R) (345, 346) and the tumor necrosis factor-α (TNF-α) receptor (347) have been reported to be activated after IR resulting in intracellular signaling cascades that lead to gene expression that mediates resistance to IR-induced cell killing (348). The epidermal growth factor (EGF) family of tyrosine kinase receptors, including EGFR (also called Erb1), Erb2, Erb3 and Erb4, are thought to mediate cell survival after IR exposure through the activation of the MAPK cascade (344, 349). IGF-1R over-expression can also mediate radio-resistance through the activation of the Mek/Erk and PI-3 kinase pathways (348). Furthermore, the TNF-α receptor is thought to mediate radio-resistance by activation of NF-κB (350). IR also increases production of the ligands for these receptors (e.g., EGF, TGF-α, IGF-1, and TNF-α) resulting in activation of the receptor and induction of downstream signaling pathways (345, 349, 351, 352). Since these signaling cascades can enhance the survival of cancer cells, their stimulation represents pathways that limit the effectiveness of radio-therapy in the treatment of a variety of

90 human malignancies that over-express these receptors. Thus, elucidating mechanisms that may be targeted by inhibitors in combination with radiotherapy should improve therapeutic efficacy. Indeed, small molecule inhibitors and antibodies to the EGFR family members are already in clinical trials (353-357), but these therapies will only be effective if EGFR is the major component of protection and resistance.

Clusterin (CLU) is a highly glycosylated secretory protein thought to provide cytoprotection after various cell stresses due to its role as a molecular chaperone. sCLU has been shown to clear and mediate the turn-over of extracellular debris (75, 76, 78, 325,

358). CLU plays a direct role as a cellular stress sensor, since a variety of cytotoxic agents (topotecan, etoposide, colcemid, nocodazole, taxol and thapsigargin) induce CLU gene and protein expression (94) at doses far below their cytotoxic effects.

CLU has also been implicated in many normal biological processes including tissue remodeling (21, 359, 360), sperm maturation (361, 362), complement inhibition

(15, 305, 363, 364), lipid transport (9, 13, 365), and pathological states ranging from

Alzheimer disease (39, 309, 366, 367), atherosclerosis (37, 307, 308, 368, 369), prion diseases (370-372), rheumatoid arthritis (45, 373), glomerulonephritis (82, 374, 375) and many cancers (51, 53, 57, 59, 376-379). Over-expression of endogenous CLU correlates with higher tumor grade and poor prognosis in prostate and breast cancer (49, 51, 58,

377). Additionally, over-expression of exogenous CLU results in resistance to paclitaxel

(62, 310), doxorubicin (87), cisplatin (63) and radiation therapy (380), while antisense

CLU expression enhances the chemosensitivity of various cell lines (73, 74). Together these data suggest that sCLU provides a survival advantage for the cancer cells that over-

91 express the endogenous protein. Thus decreasing or eliminating sCLU levels during various cancer therapies may increase the efficacy of these treatments.

Our laboratory identified CLU as an IR-induced protein/transcript (91). Secretory clusterin (sCLU), the fully processed, glycosylated form of this protein is induced by IR doses as low as 2 cGy (35, 94). Interestingly, peak CLU promoter activity, transcript and protein levels occurred 48-72h after IR exposure (94). To date, the signaling cascades leading to CLU upregulation after IR remain undetermined. Several transcription factors have been implicated in sCLU induction after various stress responses, including c-fos

(86), c-myb (87) and the TGF-β pathway (84-86), although the signaling pathways involved in the induction after IR have not been elucidated.

In this report, we show that sCLU is involved in cell survival after IR, by siRNA.

Furthermore, the activation of the Src-Raf-Mek-Erk signaling cascade, by clinically relevant doses of IR, in MCF-7 breast cancer cells is required for sCLU induction.

MAPK signaling was detected in cells within hours after IR as previously reported (344,

348, 349). However, we discovered a dramatic re-activation of MAPK signaling 24-72 h after IR. This delayed activation of MAPK was required for the late induction of the

CLU promoter, and subsequent elevations in sCLU transcript and protein. In defining the temporal and dose-response relationships of activation of this delayed signal transduction pathway, we identified a role for the insulin-like growth factor-1 receptor (IGF-1R) in the induction of sCLU after IR. Since sCLU has been shown to be a cytoprotective protein, these results may explain the radio-resistant phenotype of tumor cells that over-express the IGF receptor (345, 348, 352). A better understanding of the signal transduction

92 pathways activated by IR that stimulate sCLU levels in cells will allow for improved efficacy of radiation therapy against many human malignancies.

Experimental Procedures

Plasmids

The Src CA, Src KD, dn Erk-1, dn Erk-2 and Egr-1 plasmids were a generous gift from

Dr. Lindsey Mayo at Case Western Reserve University. The dn Mek-1 plasmid was a kind gift from Dr. Jeff Holt at Vanderbilt University (381). The 4250 CLU promoter was constructed as follows. The following primers were used to amplify the 4676 bp fragment of human CLU promoter:

5’ TGG CTA GCC TCA GAC CTT TAC CCA AGA AGA TC 3’

5’ GCT CGA GCG TTG TGG GCA CTG GGA G 3’

Genomic DNA from MCF-7 cells was used as a template. PCR was performed using

Elongase enzyme mixture (Invitrogen) with 1.8-2 mM of MgCl2. PCR conditions were:

94˚C – 45 sec; 63˚C – 30 sec; 68˚C – 5 min; 35 cycles. The 4676 bp product was gel

purified, and a 4250 bp was inserted into the SmaI site of pA3luc plasmid (a gift from Dr.

Richard G. Pestell, Georgetown University, Washington, DC). The orientation of the insert was verified by restriction digest and partial sequencing (Leskov et al., unpublished data).

93 Cell Culture and Transfections

MCF-7 breast cancer cells were grown in RPMI 1640 with 5% FBS at 37˚C in a humidified incubator with 5% CO2-95% air atmosphere. Media was replenished every 24 h on experimental plates.

PP1 was obtained from BioMol (Plymouth Meeting, PA) and U0126 was obtained from Cell Signaling Technology (Beverly, MA). AG1478 and AG1024 were obtained from Calbiochem (La Jolla, CA). EGF and IGF were obtained from Sigma or

Calbiochem respectively. EGF/IGF experiments were performed as follows. Cells were plated at a confluency of 5 X 105 cells per 10 cm2 plate. Twenty-four hours later, normal growth media was removed and replaced with media containing no serum. The next day, cells were pretreated for 1 h with either AG1478, AG1024 or DMSO and then were exposed to 5 Gy ionizing radiation with 137Cs gamma rays at a dose rate of 0.87-0.92

Gy/min, using a Shepard Mark Irradiator or treated with IGF or EGF in media containing no serum. One hour after radiation or IGF/EGF treatment, cells were replated into media containing 1% serum and were maintained in 1% serum until harvest at 48 h. AG1478 and AG1024 were removed 24 h after IR or IGF/EGF treatment.

PP1 and U0126 experiments were performed as follows. Cells were plated at a confluency of 5 X 105 cells per 10 cm2 plate. The next day, cells were pretreated for 1 h with the either PP1, U0126 or DMSO then exposed to 5 Gy. Untreated cells were mock- irradiated as described (11). In Figures 5A and 5B, media containing 20 µM PP1 was removed from the cells 24 h after IR. In all other experiments a continuous exposure of

PP1 or U0126 was used, and new PP1 or U0126 in fresh media was added to the cells every 24 h until harvest at 72 h.

94 The 1403 cell line stably expressing 1403 bp of the CLU promoter linked to the luciferase gene was generated as previously described (94). MCF-7 cells were transiently transfected using Effectene (Qiagen; Valencia, CA), according to the manufacturers directions.

Luciferase Assays

Luciferase assays with the 1403 CLU promoter were performed with the Luciferase

Assay System (Promega; Madison, WI). Cells were seeded in 6-well plates at approximately 50% confluency. Cells were treated with inhibitor and/or 5 Gy and harvested at the indicated time point in 1X reporter lysis buffer (Promega; Madison, WI).

Luciferase assays performed on MCF-7 cells transiently transfected with the 4250 CLU promoter and Src CA, Src KD, dn Mek-1 or dn Erk-1/2 plasmids were performed using the Dual Luciferase System (Promega; Madison, WI). MCF-7 cells were transfected 24 h before treatment with inhibitor and/or 5 Gy and harvested in 1X reporter lysis buffer at the indicated times. All experiments were equalized for protein using Bradford Assays

(Bio-Rad Laboratories; Hercules, CA). Each dose/time point was completed at least in triplicate and a Student’s T-Test was performed to determine statistical significance.

Western Blot Analyses and Co-Immunoprecipitations

Whole cell extracts were extracted in RIPA buffer (0.1% SDS, 0.5% deoxycholate, 1%

NP-40, 150 mM NaCl, 50 mM Tris pH 8.0) and separated on a 12% gel by SDS-PAGE western blot analyses as previously described (35). Proteins were transferred to

Immobilon-P (Millipore; Bedford, PA) and probed with antibody. Antibodies to human

95 sCLU (B5), Ku70 (C-19), P-Raf-1 (Ser338), Raf-1 (C-12), P-Mek1/2 (Ser218/222), Mek 1 (C-

18), P-Erk (E-4), Erk 1 (K-23), P-Fak (Tyr925), Egr-1 (588) and Sp1 (PEP2) were obtained from Santa Cruz (Santa Cruz, CA). IGF-1 receptor α, P-EGF receptor (Tyr1068),

EGF receptor and c-Src (Tyr416) antibodies were obtained from Cell Signaling

Technology (Beverly, MA). α-tubulin and c-Src (Tyr416) antibodies were obtained from

Calbiochem (La Jolla, CA).

EGFR was immunoprecipitated from 250 µg total protein using 2 µg antibody at

4˚C overnight. Complexed lysates and antibodies were incubated with protein G/agarose beads at 4˚C for 1 h. Complexes were washed three times with RIPA buffer, resuspended in 25 µl of SDS loading buffer, boiled for 5 minutes and separated on a 12% SDS-PAGE gel.

Quantitation was performed using NIH Image. Relative levels were determined by dividing the experimental protein levels by the total protein levels.

siRNA and Clonogenic Survival Assays

The following siRNAs against sCLU mRNA (siRNA-sCLU) or to a scrambled sequence were synthesized by Dharmacon, Inc. (Lafayette, CO): siRNA-sCLU 5'– GCG UGC AAA GAC UCC AGA AdTdT–3'

3'–dTdTCGC ACG UUU CUG AGG UCU U–5' siRNA-scrambled 5’-GCG CGC UUU GUA GGA UUC GdTdT-3’ 3’-dTdTCGC GCG AAA CAU CCU AAG C-5’

5 ug of siRNA-CLU, or scrambled siRNA was transfected into MCF-7 cells (5x105 cells/60 mm dish) using Lipofectamine Plus (Invitrogen, Inc.; Carlsbad, CA) according to the manufacturer’s instructions. Mock-transfected cells were used as a control. 48 h after

96 transfection, cells were trypsinized, and plated onto 60 mm dishes (500 cells per dish) in triplicates. Ten days later, cells were fixed and stained using crystal violet. Colonies containing >50 normal looking cells were counted.

siRNA to Egr-1 was generated by Dharmacon, Inc. (Lafayette, CO) as a Custom

SMARTpool that contained four siRNA sequences against Egr-1. MCF-7 cells were transfected with 20 µM siRNA-Egr-1, 20 µM scrambled siRNA or mock transfected using Oligofectamine following the manufacturers protocol (Invitrogen, Inc; Carlsbad,

CA). Twenty-four hours after transfection, cells were exposed to 5 Gy or mock- irradiated. Cells were harvested in 1 X reporter lysis buffer for luciferase assays or RIPA buffer for western blot analyses.

DNA Pull-Down Assays

A biotinylated full length CLU promoter was amplified from the 1403 pA3luc plasmid using the following primers ordered form Integrated DNA Technologies (Coralville, IA):

5’ GAT CCA TTC CCG ATT CCT 3’

5’ /5Bio/ AGC CAA GCT TCC TGT GCC 3’.

Nuclear extracts were harvested from DMSO, PP1 or U0126 treated MCF-7 cells that were either treated with 5 Gy or mock-irradiated as previously described (382). Three µg of the biotinylated CLU promoter were incubated with 10 µl of strepavidin beads

(Oncogene Research Products; Boston, MA) for 1 h at room temperature. The complex

was washed in binding buffer (50 mM Tris-HCL pH 7.5, 5 mM MgCl2, 2.5 mM EDTA,

2.5 mM DTT, 250 mM NaCl, 0.25 µg/µl poly (dI-dC)· poly (dI-dC) and 20% glycerol) before the addition of 100 µg nuclear extract and binding buffer up to 20 µl and

97 incubated at room temperature for 20 minutes. One ml of binding buffer was added and the complexes were incubated at 4˚C overnight. 24 h later, complexes were washed twice in binding buffer, beads were resuspended in 20 µl of SDS-PAGE loading buffer and separated on a 12% gel.

Results

siRNA to sCLU in MCF-7 Cells Causes Radio-sensitization

It has been reported that induction of sCLU by various chemotherapeutic agents provides cytoprotection for the cell (73, 74, 87). To determine if sCLU provides a protective role after IR, siRNA specific to the leader peptide in sCLU was transiently transfected into MCF-7 cells. sCLU protein levels were decreased 2-fold in MCF-7 cells using siRNA to sCLU (Figure 3.1B).

Decreased sCLU protein levels resulted in a significant increase in the lethality of cells treated with increasing doses IR, as determined by clonogenic survival assays (Figure 3.1A). In contrast, mock-transfected or scrambled siRNA (scr-siRNA) did not alter the survival of irradiated MCF-7 cells (Figure 3.1A), or sCLU protein levels (Figure 3.1B). These results illustrate the cytoprotective role of sCLU in irradiated MCF-7 cells, and suggest that alterations that decrease sCLU induction could alleviate the survival advantage after IR that sCLU provides cells.

98 A

1

0.1 * * 0.01 Mock

Surviving Fraction Scr-siRNA * SCLU-siRNA 0.001 * 0 2 4 6

Dose (Gy) *p<0.01

B Mock- scrambled sCLU transfected siRNA siRNA

sCLU

α-tubulin

Relative Levels: 1 0.9 0.5

Figure 3.1. siRNA knock-down of sCLU in MCF-7 Cells Causes Radio-sensitization.

(A) siRNA to sCLU increases radio-sensitivity of MCF-7 cells. Cells were transfected as described in “Experimental Procedures” and allowed to grow 10 days before being stained with crystal violet and counted for clonogenic survival assay.

(B) siRNA to sCLU diminishes sCLU protein expression. siRNA-sCLU was transfected into MCF-7 cells. Protein was harvested 60 h later and separated by

SDS-PAGE. Relative levels were determined as compared to mock-transfected controls.

99 Activation of MAPK by IR

It is generally presumed that the growth factor induced signaling pathways are activated and then shutdown very quickly after a given stimulus. Since sCLU is expressed late after

IR, with induction occurring 48-72 h after exposure (94), it was not understood how early signaling events could result in the late activation of this gene. To examine the temporal activation of kinases that may correlate with sCLU induction, we used western blot analyses to examine the activation kinetics of Src, Raf, and Mek in MCF-7 cells after 5

Gy (Figure 3.2).

Phospho-specific antibodies that indicated activation of these proteins, were compared to changes in total protein levels. P-Src (Tyr416), P-Raf (Ser338), P-Mek-1/2

(Ser218/222) and P-Erk-1/2 (Tyr204) were detected after IR at 0.25 h, 0.5 h, 0.5-1.0 h, and

0.25-2 h, respectively. Thus, as previously reported after high doses of IR (344), the

Src/Raf/Mek cascade was induced immediately after IR exposure. Surprisingly, however, a dramatic re-activation of the Src/Raf/Mek/Erk1/2 pathway was noted at 24-72 h after IR that appeared to correlate with sCLU protein induction in MCF-7 cells. Taken together, the data in Figure 3.2 show that the MAPK cascade was reactivated in MCF-7 cells starting at 24 h after IR in a temporal sequence that correlated with sCLU protein induction after IR.

EGFR is Not an Upstream Activator of Clusterin

Prior work demonstrated that IR can activate the epidermal growth factor receptor

(EGFR), resulting in activation of the Raf-Mek-Erk MAPK cascade (344, 349, 383). In light of this work, we examined the role of EGFR in mediating CLU expression after IR,

100 using 1403 MCF-7 breast cancer cells, as previously described (94). Serum-starved cells were pretreated with AG1478, a selective inhibitor of EGFR kinase activity, for 1 h and cells were then mock-irradiated or exposed to 5 Gy. Increasing doses of AG1478 did not affect IR-induction of the CLU promoter monitored 48 h post-IR. AG1478 did not affect basal CLU promoter activity (Figure 3.3A).

To determine if epidermal growth factor (EGF) could directly induce sCLU expression, CLU promoter activation and sCLU protein levels were examined after various doses of EGF. EGF did not induce CLU promoter activity over control levels, and various doses of AG1478 did not significantly affect CLU promoter activity before or after IR (Figure 3.3B). These results were confirmed by western blot analyses (Figure

3.3C), where sCLU was not induced by treatment with 1 µM AG1478 (Lane 4), although

EGF (1 ng/ml) induced sCLU protein levels (Lane 5) and this induction was also not blocked by treatment with AG1478 (Lane 6). We do not fully understand the mechanism of EGF induction of sCLU, but this could be a direct effect of other EGFR family members that are not inhibited by AG1478. We are currently investigating this. Since

AG1478 was able to attenuate EGFR activation (Figure 3.3D) but did not alter IR induction of sCLU, we can conclude that the EGFR does not play a role in the induction of sCLU after IR.

Inhibition of Insulin-Like Growth Factor-1 Receptor Abrogates Clusterin Induction after IR

Given that inhibition of EGFR by AG1478 had no effect on sCLU expression while EGF treatment of MCF-7 cells resulted in elevated sCLU protein levels, we examined the

101 Time (h) Post-IR 0 0.25 0.5 1 2 4 8 24 48 72 (h): P-src

src

Relative Levels 1.0 7.4 11.6 5.1 4.2 12.9 12.8 15.4 14.1 14.5 (P-Src):

P-Raf

Raf

Relative Levels 1.0 1.0 1.3 0.9 0.9 0.6 1.2 0.6 1.4 1.1 (P-Raf):

P-Mek

Mek

Relative Levels 1.0 0.6 0.4 0.5 0.7 0.6 0.8 1.8 1.6 2.5 (P-Mek):

P-Erk 1/2

Erk 1

Relative Levels 1.0 2.7 1.7 1.9 1.5 1.3 1.6 3.3 3.9 2.0 (P-Erk1/2):

SCLU (60 kDa)

Ku70

*Relative Levels 1.0 0.8 0.8 0.8 0.8 0.9 1.0 1.5 1.8 2.0 (sCLU):

Figure 3.2. Activation of MAPK by IR.

The MAPK cascade is activated in MCF-7 cells after exposure to 5 Gy. MCF-7 cells were exposed to 5 Gy and harvested for western blot analyses at the indicated time points. Western blot analyses were performed as described in “Experimental

Procedures”. Relative levels determined as irradiated samples compared to unirradiated control samples.

*Relative sCLU levels were calculated with respect to Ku70 levels.

102 A B 600 600 UT DMSO 500 5 Gy 500 AG1478 400 400

300 300 Promoter Activity Promoter Activity 200 200 CLU at 48h Post-IR CLU at 48h Post-IR 100 100

0 1

Relative 0

0 0.5 1.0 5.0 0 0.5 1.0 5.0 Relative 1 0 0.5 1.0 5.0 0 0.5 1.0 5.0 AG1478 Dose (µM) EGF Dose (ng/ml)

C IR (5 Gy) - + + - - -

AG1478 (1.0 µM) - - + + - +

EGF (1.0 ng/ml) - - - - + + sCLU Ku70 Fold Induction: 1 2 2 1 2 2

D IP: EGFR IR (5 Gy) UT IR & AG1478 AG1478 EGF EGF & AG1478 WB: P-EGFR WB: EGFR Fold Induction: 1.0 1.1 0.6 1.3 2.0 1.3

103 Figure 3.3. EGFR is Not an Upstream Activator of Clusterin.

CLU promoter activity and protein expression were not affected by treatment with the EGFR inhibitor, AG1478. MCF-7 cells or 1403 cells were placed in 0% serum,

24 h before irradiation or addition of EGF. One hour prior to irradiation or addition of growth factor, cells were treated with DMSO or the indicated concentration of AG1478. One hour after irradiation or addition of growth factor, media was removed and fresh media containing 1% serum was added to the cells.

Cells were maintained in 1% serum until harvest. One day after irradiation or addition of growth factor, AG1478 was removed. All samples were harvested 48 h after irradiation or addition of growth factor. (A) Increasing doses of AG1478 are not able to abrogate CLU promoter activity. (B) Increasing amounts of EGF, with and without AG1478, have no effect on CLU promoter activity. (C) AG1478 is not able to abrogate sCLU protein induction 48 h after exposure to 5 Gy. (D) EGFR co- immunoprecipitation shows that EGFR is phosphorylated on Tyr1068 after exposure

to 5 Gy or addition of 1.0 ng/ml EGF, and that this phosphorylation can be blocked

by pretreatment with 1µM AG1478.

104 possible involvement of insulin-like growth factor receptor 1 (IGF-1R) in IR-mediated

CLU induction. The IGF-1R is a potent activator of the Raf-Mek-Erk MAPK cascade

(384-386), however, involvement of IGF-1R after IR has not been explored in detail.

Serum-starved 1403 cells were pretreated with increasing doses of AG1024, an IGF-1R kinase inhibitor, and either mock-irradiated or exposed to 5 Gy (Figure 3.4A). Serum- starvation of MCF-7 cells increased the basal levels of sCLU promoter activity.

Treatment of cells with AG1024 resulted in significantly reduced CLU promoter activity compared to DMSO controls (Figure 3.4A). Interestingly, induction of the 1403 CLU promoter was not observed with increasing doses of IGF. Furthermore, when cells were pretreated with 2.5 µM AG1024, the observed elevated basal level of CLU promoter activity was abrogated (Figure 3.4B), suggesting a direct role for IGF-1R signaling in

CLU gene expression. These results were then confirmed at the protein level by western blot analyses (Figure 3.4C). As previously noted, sCLU was induced over basal levels

(Lane 1) by 5 Gy (Lane 2), and this induction was blocked by pre-treatment with 2.5 µM

AG1024 (Lane 3, Figure 3.4C). Treatment of MCF-7 cells with AG1024 alone had little affect on basal sCLU protein levels (Lane 4). IGF (10 nM) had no effect on sCLU protein expression in MCF-7 cells (Lane 5), but basal levels were reduced by pre-treatment with

AG1024 (Lane 6). Interestingly, we noted that IGF-1R was slightly induced in MCF-7

105 A B 500 800 450 UT 700 DMSO 400 5 Gy 350 600 AG1024 300 500 250 400 200 *

Promoter Activity * 300 150 * Promoter Activity 100 200 CLU at 48h Post-IR CLU at 48h Post-IR * * * 50 100 * 0 1 0 0 1.0 2.5 5.0 0 1.0 2.5 5.0 1 0 5 10 20 0 5 10 20 Relative AG1024 Dose (µM) Relative *p<0.04 IGF Dose (nM) *p<0.05

C IR (5 Gy ) - + + - - -

AG1024 (1.0 µM) - - + + - +

IGF (10 ng/ml) - - - - + +

sCLU

Ku70

Relative Induction: 1.0 1.6 1.1 1.1 1.0 0.8

D IR (5 Gy ) - + + - - -

AG1024 (1.0 µM) - - + + - +

IGF (10 ng/ml) - - - - + +

IGFR

Ku70

Relative Induction: 1.0 1.2 0.6 0.9 0.8 1.0

0.095

0.090 0.085 * 0.080 * *

0.075 IGF (ng/ml) 0.070 whole serum 0.065 serum starved

0.060 0 24 48 72 Time (h) Post-IR (5 Gy) *p<0.01

106 Figure 3.4. Inhibition of Insulin-Like Growth Factor-1 Receptor Abrogates

Clusterin Induction after IR. CLU promoter activation and protein expression were abrogated by treatment with the IGF-1R inhibitor, AG1024. MCF-7 cells or 1403 cells were placed in 0% serum, 24 h before irradiation or addition of IGF. One hour prior to irradiation or addition of growth factor, cells were treated with

DMSO or the indicated concentration of AG1024. One hour after irradiation or addition of growth factor, media was removed and fresh media containing 1% serum was added to the cells. Cells were maintained in 1% serum until harvest.

One day after irradiation or addition of growth factor, AG1024 was removed. All samples were harvested 48 h after irradiation or addition of growth factor. (A)

Basal and inducible CLU promoter activation is abrogated by pretreatment with increasing amounts of AG1024. (B) AG1024 decreases CLU promoter activity in serum-starved MCF-7 cells treated with increasing amounts of IGF. (C) Western blot analyses show that AG1024 is capable of abrogating sCLU protein induction after exposure to 5 Gy. (D) IGF-1R protein levels are upregulated in MCF-7 cells stimulated with either 5 Gy or 10 ng/ml IGF, and this up-regulation can be blocked by pretreatment with 2.5 µM AG1024. (E) An ELISA to human IGF-1 demonstrates that IGF-1 production and secretion is stimulated approximately 20% in MCF-7 cells 24-72 h after exposure to 5 Gy. *p value signifies statistical difference between irradiated samples and mock-irradiated samples.

107 cells after 5 Gy (Figure 3.4D). Treatment of cells with AG1024 and 5 Gy dramatically decreased IGF-1R levels (Lane 3). Addition of AG1024 also lowered the basal levels of

IGF-1R in unirradiated, serum starved cells, suggesting a possible autocrine feedback loop induced by IR, where irradiated cells up-regulate IGF-1R.

Since the receptor for IGF was elevated after IR, we investigated whether IGF-1 production was also elevated after exposure of cells to 5 Gy. Media collected from serum-starved MCF-7 cells at 0, 24, 48 and 72 h post-IR showed increased production of approximately 20% in irradiated MCF-7 cells at 24 h after IR, with sustained increases noted by 72 h post-IR (Figure 3.4E). These data suggest that MCF-7 cells have high basal levels of IGF-1 as reported (387), and that after exposure to 5 Gy, both he receptor and the ligand increase at later times (24-72 h) forming a putative autocrine feedback loop.

c-Src is an Upstream Activator of sCLU Induction After IR

It is known that c-Src, but not other Src family members, can activate and be activated by

IGF-1R (388-391). To determine whether c-Src activation was required for sCLU induction after IR, we used PP1, a selective inhibitor of Src family kinases. MCF-7 1403 cells were pretreated for 1 h with 20 µM PP1 and exposed to increasing doses of IR. PP1 was removed 24 h post-exposure, and luciferase activity was assessed at 72 h. PP1 abrogated induction of the sCLU promoter at low doses of IR, inhibiting sCLU induction by 50% at higher doses of IR (Figure 3.5A). As previously noted (35, 94), sCLU was modestly induced by 10 cGy in DMSO-treated cells, and this induction was blocked by

108 20 µM PP1 treatment (Figure 3.5B). PP1 inhibited IR-induced sCLU expression, with only minor inhibition noted after 2.5 Gy (Figure 3.5B), potentially due to the metabolism of PP1 by 72 h.

To address the potential loss of PP1 inhibition after a 24 h treatment period, we determined the lowest dose of PP1, given as a continuous 72 h treatment that could inhibit CLU promoter activity after 5 Gy. PP1 (5 µM) significantly decreased the basal activity of the CLU promoter and effectively blocked IR-activated CLU promoter activity

(Figure 3.5C). sCLU protein levels were also concomitantly decreased (Figure 3.5D). To confirm that the PP1 inhibitor was functional, we examined a known substrate of c-Src,

Fak. Fak is phosphorylated at Tyr925 by c-Src (391). Phosphorylation of Fak decreased with increasing doses of PP1 (Figure 3.5E). Thus, PP1 effectively blocked c-Src activity at doses (1 µM and 5 µM) that corresponded to doses used to block sCLU induction.

To confirm a role for c-Src as an upstream regulator of sCLU induction after IR, we transiently over-expressed increasing amounts of a constitutively active Src plasmid

(Y529F) or a kinase dead Src plasmid (K297R) in 1403 cells. Transfected 1403 cells were mock-irradiated or treated with 5 Gy and harvested at 72 h for CLU promoter activity. Titration of increasing amounts of constitutively active Src increased basal and

IR-inducible CLU promoter activity (lanes 1-6, Figure 3.5F). In contrast, titration of increasing amounts of kinase dead Src (Src KD) repressed IR-induced CLU promoter activity. Collectively, these data using PP1, a selective c-Src inhibitor, and expression of constitutively active or kinase dead Src, demonstrate the novel observation that c-Src is required for sCLU induction after IR.

109 A

1000 900 B 800 DMSO PP1 (20 M) 700 µ 600 IR (Gy): 0 0.1 0.25 0.5 1.0 2.5 0 0.1 0.25 0.5 1.0 2.5 500 * 400 sCLU Promoter Activity 300 200 Ku70 CLU at 72h Post-IR 100 0 IR (Gy): 0 0.1 0.25 0.5 1.0 2.5 5.0 1 0 0.1 0.25 0.5 1.0 2.5 5.0 PP1 (20 M) Relative DMSO µ *p<0.03

C 2.E+05 D 2.E+05 UT 2.E+05 1.E+05 5 Gy UT IR (5 Gy) 1.E+05 * 1.E+05 PP1 (µM): 0 1 5 10 0 1 5 10 8.E+04 * * 10 Promoter Activity * 6.E+04 sCLU

4.E+04 CLU at 72h Post-IR * * * 2.E+04 Ku70 0.E+00 1 0 1 5 10 15 0 1 5 10 15

Relative PP1 Dose (µM) *p<0.05

F E 800 700 * UT 600 * 5 Gy UT IR (5 Gy) 500

400 0 0 1 5 10 * * PP1 (µM): 300 * 200 Promoter Activity

P-Fak 100

CLU 0 at 72h Post-IR 1 1 2 3 4 5 6 7 8 9 10 11 Ku70 12 DNA (µg): Src CA Src KD Relative *p<0.05

110 Figure 5. c-Src is an Upstream Activator of sCLU Induction After IR

A 24 h exposure of MCF-7 cells to 20 µM PP1 can abrogate CLU promoter activation and sCLU protein expression after 5 Gy. (A) 20 µM PP1 can inhibit CLU promoter activation after increasing doses of IR. 1403 cells were pretreated with 20

µM PP1 one hour before irradiation. Media was removed, cells were washed with

PBS and fresh media without inhibitor was added 24 h after IR. (B) 20 µM PP1 can inhibit sCLU protein induction by IR doses as low as 0.1 Gy. MCF cells were treated as described above. (C) Basal and IR-inducible activity levels of the CLU promoter are inhibited by continuous treatment (over 72 h) of PP1. Increasing doses of PP1 were added to 1403 cells one hour before irradiation. Media with PP1 was replenished every 24 h. (D) sCLU protein induction by IR is abrogated in cells treated continuously with PP1. Increasing doses of PP1 were added to MCF-7 cells one hour before irradiation. Media with PP1 was replenished every 24 h.

(E) PP1 is able to inhibit phosphorylation of the Src downstream substrate, Fak, at doses that abrogate sCLU protein induction after 5 Gy. MCF-7 cells were pretreated for 1 h with increasing doses of PP1 and harvested for western 72 h after

IR. (F) MCF-7 cells transiently transfected with a plasmid expressing a constitutively active Src protein (Src CA) (Lanes 1-6) show an increase in basal and inducible promoter activity and cells transiently transfected with a kinase dead Src protein (Src KD) (Lanes 7-12) show inhibition of IR-inducible CLU promoter activity. Lanes 1, 4, 7 and 10 were transfected with vector only DNA. Lanes 2, 5, 8 and 11 were transfected with 0.2 µg of the appropriate DNA and lanes 3, 6, 9 and 12 were transfected with 0.3 µg of the appropriate DNA. All transfections contained a

111 total of 3 µg DNA. Cells were transfected 24 h prior to irradiation and harvested 72 h after IR.

Activation of the MAPK cascade is required for sCLU Induction After IR

Src phosphorylates the downstream signaling kinase, Raf, which in turn activates Mek-1 through phosphorylation (392). To examine the requirement of Mek-1 for sCLU activation, we used U0126, a selective Mek-1 kinase inhibitor, after IR. MCF-7 1403 cells were pretreated for 1h with increasing concentrations of U0126 and then exposed to

5 Gy. IR induction of the sCLU promoter was completely abrogated by 1µM U0126

(Figure 3.6A). Consistent with loss of gene induction, 1 µM U0126 completely abolished IR-induced sCLU protein in MCF-7 cells (Figure 3.6B).

The Erk 1/2 kinases are downstream substrates for Mek-1 (349). To demonstrate the involvement of Erks in CLU gene induction, we over-expressed dominant-negative

Erk-1 (dn Erk-1) or dominant-negative Erk-2 (dn Erk-2) in 1403 cells. After transfection, we mock-irradiated or exposed cells to 5 Gy and luciferase activity was examined 72 h later (Figure 3.6C). Both dn Erk-1 and dn Erk-2 completely suppressed CLU promoter activity after 5 Gy.

Log-phase MCF-7 cells were co-transfected with a 4250 bp CLU promoter- luciferase reporter and Src CA, Src KD, dominant-negative Mek-1 (dn Mek-1 K97A) or dominant-negative Erk-1 (dn Erk-1) (Figure 3.6D). In response to 5 Gy, the 4250 CLU promoter was induced approximately 4-fold in MCF-7 cells. Over-expression of constitutively active Src resulted in the elevation of the CLU promoter basal level 4-fold

112 A 1.E+04 B 1.E+04 UT * UT IR UT IR (5 Gy) 1.E+04 5 Gy 8.E+03 U0126 (µM): 0 0 5 10 1 5 10

Promoter Activity 20 6.E+03 sCLU

CLU 4.E+03 at 72h Post-IR Ku70 2.E+03

0.E+00 1 0 1 5 0 1 5 Relative U0126 Dose (µM) *p<0.05

C

250 UT 200 5 Gy E 150 Src Src VO CA KD Promoter Activity 100 * * * IR (5 Gy): - + - + - + CLU at 72h Post-IR WB: Src 50 WB: Ku70

Relative 0 dn 1 dn dn VO Erk-1 Erk-2 Erk-1/2 dn VO *p<0.02 Mek-1 IR (5 Gy): - + - + D WB: Mek-1 4.E+04 * WB: Ku70 4.E+04 UT 3.E+04 5 Gy 3.E+04 * dn VO Erk-1 Promoter Activity 2.E+04 2.E+04 IR (5 Gy): - + - + CLU at 72h Post-IR 1.E+04 WB: HA 5.E+03 * * * NB 0.E+00 1 Relative Src Src dn dn VO CA KD Mek-1 Erk-1 *p<0.05

113 Figure 3.6. Activation of the MAPK cascade is required for sCLU Induction After

IR. The Mek-Erk-Egr-1 pathway is required for sCLU induction after 5 Gy.

(A) The Mek-1 inhibitor, U0126, decreases IR-inducible CLU promoter activity to basal levels. 1403 cells were pretreated with increasing amounts of U)126 1 h prior to irradiation. Media with inhibitor was replenished every 24 h until cells were harvested 72 h after IR. (B) U0126 abrogates sCLU protein expression in

MCF-7 cells exposed to 5 Gy. MCF-7 cells were pretreated with increasing amounts of U)126 1 h prior to irradiation. Media with inhibitor was replenished every 24 h until cells were harvested 72 h after IR. (C) Transient transfection of dn Erk-1 or dn

Erk-2 into MCF-7 cells is able to diminish CLU promoter activity to basal levels after 5 Gy. (D) MCF-7 cells were transiently co-transfected with 4250 bp of the

CLU promoter in the pA3luc vector with either vector only (VO), constitutively active Src (Src CA), kinase dead Src (Src KD), dn Mek-1 or dn Erk-1. (E) Western blots confirming over-expression of Src, Mek-1 or Erk-1 in MCF-7 cells.

114 compared to control. Upon treatment with IR, a further increase in CLU promoter activity was observed in Src CA transfected cells. Over-expression of kinase dead Src, dominant-negative Mek- 1 and dominant-negative Erk-1 abrogated IR-induction of this promoter. Western blot analyses confirmed over-expression of the corresponding proteins from the plasmids used above (Figure 3.6E). These observations demonstrate that the

Src/Mek/Erk pathway was required for CLU promoter induction and sCLU protein expression in response to IR. Collectively, these data demonstrate the involvement of the

MAPK signaling cascade in sCLU induction in MCF-7 cells after IR.

The Egr-1 Transcription Factor can Bind to the CLU Promoter and is Required for

Promoter Activation After IR

There are several transcription factors that are known to be activated by the IGF-

1R/Src,Raf/Mek/Erk signaling cascade. Examination of the CLU promoter identified three potential Egr-1 binding sites (at –450 bp, -918 bp and –932 bp relative from ATG start site). Egr-1 is a known downstream target of the MAPK cascade (383, 393-396), and previously shown to be activated by high doses of IR (208, 283, 284, 286). Over- expression of Egr-1 increased CLU promoter activity~2-fold over basal levels, and an ~3- fold increase was noted in MCF-7 cells after exposure to 5 Gy (Figure 3.7A). Western blot analyses confirmed the over-expression of the Egr-1 protein in these cells (Figure

3.7B). This suggested to us that the Egr-1 protein is an important transcription factor that mediates induction of CLU gene expression after IR.

DNA pull-down assays were performed to determine if Egr-1 could bind to the

CLU promoter and if IR exposure of MCF-7 cells enhanced Egr-1 binding. The 1403

115 region of the promoter was amplified using a biotin-labeled primer complementary to the

3’ end of the CLU promoter. The biotinylated promoter fragment was bound to strepavidin beads, incubated with 100 ug of nuclear extract from MCF-7 cells treated with DMSO, 5 µM PP1 or 5 µM U0126 over 72 h (Figure 3.7C). After incubation, complexed DNA and proteins were prepared for western blot analyses. Egr-1 DNA binding activity was minimal in control mock-irradiated cells. After IR, however, an increase of Egr-1 binding was noted beginning at 4 h, with a robust increase in binding observed at 24 h, with sustained levels observed through 72 h post-irradiation. In contrast, IR-treated MCF-7 cells exposed to either 5 µM PP1 or 5 µM U0126 showed no increase in Egr-1 DNA binding to the CLU promoter DNA. Nuclear extracts (10% of input) were separated by SDS-PAGE and probed for PCNA as a loading control.

Since we were able to show that the Src/Mek/Erk pathway was involved in the induction of sCLU, we next determined if IGF-1R or EGFR inhibition could block Egr-1 binding to the CLU promoter. DNA pull-down assays were used to analyze Egr-1 binding to the CLU promoter in MCF-7 cells that were treated with DMSO, 1 µM

AG1024 (IGF-1R inhibitor) or 1 µM AG1478 (EGFR inhibitor) (Figure 3.7D). Exposure of IR-treated cells with AG1478 did not affect Egr-1 DNA binding activity to the CLU promoter. Consistent with our observations of IGF-1R mediating sCLU induction after

IR, Egr-1 DNA binding activity was blocked with AG1024. Interestingly, the basal Egr-

1 DNA binding to this promoter was increased in cells pretreated with either inhibitor compared to DMSO treated cells, which may be a result of cross-talk between the receptors. Competition assays using a non-biotin labeled CLU promoter (0.1 and 5 µg) showed that Egr-1 DNA binding activity was inhibited with increased concentrations of

116 unlabeled promoter, demonstrating specificity of Egr-1 DNA binding to the CLU promoter (Figure 3.7E).

To demonstrate the requirement of Egr-1 for sCLU induction, we used Egr-1 specific siRNA (Figures 3.7F & G). Consistent with previous results, IR exposure of mock-transfected 1403 cells or 1403 cells transfected with a scrambled siRNA (scr- siRNA) resulted in a 5-6-fold induction of the 1403 CLU promoter (Figure 3.7F). When control 1403 cells were transfected with 20 nM Egr-1 siRNA, the CLU promoter was not activated after 5 Gy. Efficacy of siRNA Egr-1 was confirmed by dramatic decreases in

Egr-1 protein levels (Figure 3.7G). Thus, decreased Egr-1 protein levels abrogated induction of CLU promoter activity at 72 h observed in 1403 cells after IR.

Discussion

Our studies revealed that sCLU plays a role in cell survival after exposure to IR

(Fig. 3.1). To further understand cellular responses to IR, we investigated the signaling pathways required for the IR-induction of sCLU. In this study, we report that IGF-1R activation after IR was required for sCLU induction. We demonstrated that IR-induced sCLU was dependent on a novel reactivation of the Src/Raf/Mek/Erk signaling cascade

24-72 h after IR exposure, and that this signal culminated in the activation of the Egr-1 transcription factor (model, Fig. 3.7H).

EGFR is over-expressed or constitutively activated in many types of tumors including colorectal, breast, pancreatic and ovarian cancers (397) and is known to be a mediator of radio-resistance in several tumor types including glioblastoma multiforme and breast cancer cells through the activation of Erk-1/2 (398, 399). As a result, many

117 A

7.E+04 * B 6.E+04 UT 5 Gy 5.E+04 VO Egr-1 * 4.E+04 IR (5 Gy): - + - + WB: Egr-1 3.E+04 Promoter Activity NB 2.E+04 CLU at 72h Post-IR

1.E+04

0.E+00 1

Relative VO Egr-1 *p<0.05

C DMSO PP1 (5 µM) U0126 (5 µM) Time Post-IR (h): 0 1 4 24 48 72 0 1 4 24 48 72 0 1 4 24 48 72

Egr-1

PCNA

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 Fold 1818 18 Induction 1 1 2 6 3 7 1 2 1 1 1 1 1 2 1 1 1 0.4 :

1 µM 1 µM DMSO AG1024 AG1478 D UT IR UT IR UT IR

Egr-1

PCNA

1 2 3 4 5 6 Fold Induction: 1 2 1 0.5 1 1

E UT IR (5 Gy)

Egr-1

Non-biotinylated 1403 promoter (µg):

118 F 700 UT G 600 5 Gy Mock- Scrambled Egr-1 transfecte siRNA siRNA 500 d IR (5 - + - + - + Gy): 400 Egr-1

300 Ku70 Promoter Activity * Fold 200 Induction 1.0 1.8 1.1 1.9 1.2 0.9 CLU at 72h Post-IR : 100

0 1

Relative Scrambled Egr-1 Mock- *p<0.01 siRNA siRNA transfected (20 nM) (20 nM)

H

IR IGF-1R

IRS-1 SOS Src Grb2 Ras Raf

Mek-1

Erk-2 Erk-1

Erk-2 Erk-1

Egr-1 CLU promoter

119 Figure 3.7. The Egr-1 Transcription Factor can Bind to the CLU Promoter and is

Required for Promoter Induction After IR . Egr-1 binds to the CLU promoter and siRNA to Egr-1 abolishes CLU promoter induction by 5 Gy. (A) MCF-7 cells were transiently co-transfected with the 4250 CLU promoter and vector only (VO) or

Egr-1, 24 h prior to irradiation. Control cells were mock-irradiated (untreated,

UT). Luciferase assays were used to monitor CLU promoter activity 72 h after IR.

(B) Western blot confirming Egr-1 over-expression in MCF-7 cells. (C) DNA pull- down assay demonstrating the ability of Egr-1 to bind to the 1403 CLU promoter

(lanes 1-6) and the abrogation of this binding when cells are pretreated with 5 µM

PP1 (lanes7-12) or 5 µM U0126 (lanes 13-18) before 5 Gy. (D) The IGF-1R inhibitor, AG1024 (1 µM) diminished Egr-1 binding to the 1403 CLU promoter after exposure to 5 Gy as compared to DMSO treated cells. The EGFR inhibitor,

AG1478 (1µM) had no effect on Egr-1 binding to the 1403 CLU promoter after exposure to 5 Gy. (E) Egr-1 binding to the biotin-labeled 1403 CLU promoter can be competed away by increasing amounts (0.0, 0.1 or 5.0 µg) of non-biotin labeled

1403 CLU promoter.

(F) siRNA to Egr-1 lowers 1403 CLU promoter basal activity and IR-inducible activity as compared to mock-transfected cells or cells transfected with a scrambled siRNA. (G) Western blot analyses showing diminished Egr-1 protein levels in MCF-

7 cells after transfection with siRNA to Egr-1. Transfection with a scrambled siRNA had no effect on Egr-1 levels in MCF-7 cells. (H) Model depicting IR activation of

120 IGF-1R and the Src/Raf/Mek/Erk cascade culminating in the activation of Egr-1 and CLU promoter induction.

therapies have been developed that specifically target EGFR including monoclonal antibody therapies and small molecule inhibitors that specifically target the kinase domain (355). Interestingly, the selective EGFR inhibitor, AG1478, did not block CLU promoter induction or regulate sCLU protein levels after 5 Gy (Fig. 3.3), nor did it affect

EGF-stimulated sCLU protein expression, possibly due to the involvement of other

EGFR family members.

IGF-1R is another membrane receptor shown to be upregulated after IR. IGF-1R activation results in mitogenic growth and cell survival (384), and Gooch et al., demonstrated that treatment of cells with IGF-1 could prevent doxorubicin and taxol induced apoptosis (400). Using AG1024, a selective inhibitor of IGF-1R, we were able to block sCLU induction after IR (Fig. 3.4), demonstrating the requirement for IGF-1R activation for IR-induced sCLU expression. In a recent report, it was shown that

AG1024 treatment of MCF-7 cells enhanced cell death after IR exposure (401). Our data strongly suggests that activation of IGF-1R may mediate cell survival effects through the downstream induction of sCLU.

MCF-7 cells produce and secrete IGF-1 under serum-free conditions (387). In addition, IGF-1R is often over-expressed in breast cancer (402). It was shown previously that peripheral lymph node stromal cells produce and secrete EGF and IGF-1, which can increase the growth of breast cancer cells (403). It is possible that EGF and IGF secretion by lymph nodes can induce the tumorigenesis of neighboring breast tissue,

121 especially cells that have upregulated expression of EGFR or IGF-1R, through a paracrine mechanism. It is interesting to note that both EGF and IGF-1 were able to induce sCLU expression (Fig. 3.3C). Additionally, serum starvation increased the basal activity of the CLU promoter compared to cells grown in whole serum, suggesting a possible autocrine feedback loop induced by IR, where irradiated cells not only up- regulate IGF-1R, but also presumably increase production of the ligand, IGF-1.

Consistent with a previous report (384), we demonstrate induction of IGF-1R after IR

(Fig. 3.4D), as well as a 20% increase in secretion of IGF-1, 24-72 h post-IR (Fig. 3.4E).

Importantly, the induction of IGF-1R and its ligand provide a plausible explanation for the late induction of CLU after IR, as well as the increase in basal promoter activity after serum-starvation.

We show, for the first time, that the Src-Raf-Mek-Erk-1/2 pathway is required for

IR-induced sCLU activation. This cascade culminates in the activation of the Egr-1 transcription factor. In addition, we demonstrate a novel re-activation of the MAPK cascade after IR that correlates with the temporal activation of sCLU. The physiological relevance of this biphasic activation of MAPK is unknown. EGFR may be required for the initial induction of MAPK after IR, and inhibitors to EGFR are known to potentiate the cytotoxicity of radiation therapy (404). In addition, Lu et al., demonstrated that increased IGF-1R production in MCF-7 cells caused increased resistance to Herceptin (a monoclonal antibody to the Her2/neu receptor) induced cell death (405), suggesting a role for IGF-1R signaling in the development of resistance to this type of therapy. Our data suggests that tumor cells that survive the initial phase of treatment may develop resistance to EGFR inhibitors, as a result of the late induction of MAPK signaling

122 through IGF-R1 upregulation, and potentially the upregulation of sCLU. This suggests that the current antibody and small molecule therapies used to treat EGFR positive tumors may be optimized by the addition of inhibitors to the IGF-1R pathway.

sCLU has been shown to provide cytoprotection against doxorubicin, taxol and cisplatin in the treatment of cancer cells (63, 73, 74, 87). In this report, we show that sCLU provides cytoprotection in MCF-7 cells after IR exposure (Fig. 3.1). sCLU expression has been found to be elevated in many types of tumors, including prostate, colorectal and breast cancer (49, 51, 58). In a recent paper by Chen et al., it was shown that CLU message and protein were elevated in intestinal tumors derived from mice containing the Apc min (multiple intestinal neoplasia) mutation (59). IGF-1R and IGF-1 production are also elevated in many tumor types. Gleave et al., in a recent review, suggest using CLU and insulin-like growth factor binding proteins (IGFBPs) as targets for antisense therapy against prostate cancer (73), although they did not mention a possible connection between IGF-1R signaling and the induction of sCLU. Our data strongly suggests this connection.

It is intriguing to speculate a possible role for sCLU in bystander effects, after radiation or chemotherapeutic therapies. An increased production and secretion of sCLU by tumor cells into the lymph or vasculature system may provide a survival effect for neighboring or metastatic cancer cells. Additionally, secretion of IGF-1 by normal or tumor cells may provide a means for sCLU upregulation.

An increased understanding of the signaling pathways activated after IR will allow us to better determine which tumors will respond well to this type of therapy and which tumors will be refractive to radio-therapy. This is especially true for the delayed

123 cellular responses after IR that could have effects on carcinogenesis and cancer therapy.

Since sCLU provides resistance to IR, screening tumors for sCLU over-expression may provide a way of determining the efficacy of radio-therapy. Additionally, inhibitors to proteins in the signaling pathway leading to sCLU induction (i.e., IGF-1R/c-

Src/Raf/Mek/Erk pathway inhibitors) may be useful in treating human malignancies on their own and in combination with radiation.

124 Chapter 4: Conclusions and Future Directions

Radiotherapy is used to treat many types of cancers, but many tumors are refractory to this type of therapy or develop radio-resistance over time. A better understanding of the cellular responses to IR will allow us to improve the efficacy of this treatment. Our laboratory discovered that sCLU is induced by IR doses as low as 2 cGy

(2 rads), suggesting a role for sCLU in the cellular responses to ionizing radiation.

Understanding the processes involved in the induction of sCLU by IR will allow us to enhance the efficacy of radio-therapy and potentially screen for tumors that may be refractory to this type of treatment. Such a screen would allow an opportunity for molecular intervention to overcome these resistance mechanisms. The work described in this thesis characterized and elucidated the signal transduction pathways that regulate sCLU expression after IR exposure.

sCLU is thought to be a general stress responder, acting in a cytoprotective manner. One cytoprotective function of sCLU could be related to its ability to act as a molecular chaperone, preventing the precipitation of improperly folded proteins and aiding clearance of extracellular debris (Chapter 1). The experiments demonstrating sCLU’s ability to act in a manner similar to small heat shock proteins were performed in vitro, looking at the effects of the addition of exogenous CLU on various proteins after heat shock or (75, 406). It is intriguing to speculate on what role sCLU is performing as a secreted chaperone, since sCLU induction does not occur until several days after stress. It may be that sCLU is secreted from cells that survive cytotoxic stress,

125 and is needed to clear cellular debris from dying cells in order to “turn over” factors that could be potentially cytotoxic or inflammatory.

We found that CLU is transcriptionally repressed by the p53 tumor suppressor protein (Chapter 2). The p53 protein is stabilized in response to genotoxic stress and acts as a transcription factor for genes resulting in either cell cycle arrest or apoptosis. p53 can also act as a repressor of transcription, although the exact mechanism of repression varies. Several lines of evidence suggest that sCLU is transcriptionally repressed by p53:

(A) HCT116 colon cancer cells that are p53 null show a dramatic induction of sCLU after

IR as compared to cells that contain wild-type p53; (B) MCF-7 cells that contain the

HPV-16 E6 protein have high basal levels of sCLU as compared to cells without E6, and

(C) sCLU is induced after IR exposure in RKO cells that contain the HPV-16 E6 protein, but not in RKO parental cells that contain functional p53.

It has been suggested that sCLU is regulated through the cell cycle, with peak

induction seen as cells exit the G0 quiescent phase. Bettuzzi et al., demonstrated that

CLU message peaked in serum-starved normal human dermal fibroblasts as they entered

G0 (328). We examined sCLU expression during different phases of the cell cycle to account for the possibility that the transcriptional repression of sCLU by p53 was due to position in the cell cycle. Isogenic HCT116 cells that differ only in their p53 or p21 status were serum-starved and confluence arrested with the result that approximately 80%

of the cells arrested in G0. We noted little change in sCLU status in any of these cell lines as the cells progressed through the cell cycle, suggesting that repression of sCLU in cells with functional p53 protein was dependent on p53 and not on cell cycle phase.

126 The mechanism of CLU repression by p53 remains to be elucidated. There are several proposed models on how p53 can act as a transcriptional repressor (407). In the first model, p53 binds to its putative DNA binding sequence and sterically inhibits the binding of transcription factors required for induction. Upregulation of α-fetoprotein requires the binding of hepatic nuclear factor 3 (HNF-3) to its promoter region (333). A p53 binding site overlaps the HNF-3 binding site, thus when p53 is bound, HNF-3 is displaced. This model was also proposed to account for repression of Bcl-2 (332) and

HBV (334) genes by p53. In the second model, p53 binds and sequesters transcription factors required for transcriptional upregulation. For example, p53 can directly bind several transcription factors including Sp-1(335, 336), Ap-1 (337), NF-Y (338, 339),

Brn-3a (332) and C/EBPβ (340). In this regard, the similarities between Sp1 and Egr-1 transcription factors may be important and the possibility of p53 binding to, and repressing Egr-1 should be explored. p53 can also bind the TATA binding protein (TBP) in vitro and inhibit transcription by disrupting formation of the TFIID complex (341).

Additionally, p53 can repress the Map4 (408), α-tubulin (409) and survivin (410) genes through its interactions with the chromatin remodeling machinery. The ability of p53 to suppress survivin is of particular interest since as with sCLU, survivin is a potent pro- survival factor induced after stress. p53 associates with histone deacetylaces (HDACs) via an interaction with mSina and this interaction can be stimulated after DNA damage

(407). Alternatively, Johnson et al. have proposed a novel putative DNA binding sequence for p53 that is strictly involved in transcriptional repression (342).

Recently, our laboratory cloned a larger fragment of the CLU promoter (4250 bp) that contains a consensus p53 binding site. Using DNA pull down assays with a

127 biotinylated 4250 promoter, we have shown that p53 can bind to this site, and binding is disrupted 24-72 h after IR, which corresponds to the temporal induction of sCLU

(Leskov et al., unpublished data). This could provide a possible explanation for our data in Chapter 2, where p53 and sCLU protein levels are simultaneously high 48 h after IR in

MCF-7 cells. The p53 detected in these assays also shows a larger band that may correspond to monoubiquitinated p53. Grossman et al., demonstrate that MDM2 monoubiquitinates p53 and that p300 is required for further p53 polyubiquitination in vivo (411). It is interesting to speculate that MDM2 may be monoubiquitinating p53, which would destabilize p53 DNA binding without changing total p53 protein levels, resulting in the induction of sCLU. One could also speculate that p53 represses sCLU early after IR, due to its interaction with HDACs. After DNA repair is completed, p53 repression may be relieved via an MDM2 dependent mechanism, resulting in later elevated sCLU levels.

Our studies revealed that sCLU plays a role in cell survival after exposure to IR

(Chapter 3). To further understand cellular responses to IR, we investigated the signaling pathways required for the IR-induction of sCLU. We report that IGF-1R activation after

IR was required for sCLU induction. We demonstrated that IR-induced sCLU was dependent on a novel reactivation of the Src/Raf/Mek/Erk signaling cascade 24-72 h after

IR exposure, and that this signal culminated in the activation of the Egr-1 transcription factor (model, Fig. 3.7).

EGFR is over-expressed or constitutively activated in many types of tumors including colorectal, breast, pancreatic and ovarian cancers (397) and is known to be a mediator of radio-resistance in several tumor types including glioblastoma multiforme

128 and breast cancer cells through the activation of Erk-1/2 (398, 399). As a result, many therapies have been developed that specifically target EGFR including monoclonal antibody therapies and small molecule inhibitors that specifically target the kinase domain (355). Interestingly, the selective EGFR inhibitor, AG1478, did not block CLU promoter induction or regulate sCLU protein levels after 5 Gy (Fig. 3), nor did it affect

EGF-stimulated sCLU protein expression, possibly due to the involvement of other

EGFR family members.

IGF-1R is another membrane receptor shown to be upregulated after IR. IGF-1R activation results in mitogenic growth and cell survival (384), and Gooch et al., demonstrated that treatment of cells with IGF-1 could prevent doxorubicin and taxol induced apoptosis (400). Using AG1024, a selective inhibitor of IGF-1R, we were able to block sCLU induction after IR (Fig. 4), demonstrating the requirement for IGF-1R activation for IR-induced sCLU expression. In a recent report, it was shown that

AG1024 treatment of MCF-7 cells enhanced cell death after IR exposure (401). Our data strongly suggest that activation of IGF-1R may mediate cell survival effects through the downstream induction of sCLU.

MCF-7 cells produce and secrete IGF-1 under serum-free conditions (387). In addition, IGF-1R is often over-expressed in breast cancer (402). It was shown previously that peripheral lymph node stromal cells produce and secrete EGF and IGF-1, which can increase the growth of breast cancer cells (403). It is possible that EGF and IGF secretion by lymph nodes can induce the tumorigenesis of neighboring breast tissue, especially cells that have upregulated expression of EGFR or IGF-1R, through a paracrine mechanism. It is interesting to note that both EGF and IGF-1 were able to

129 induce sCLU expression (Fig. 3C). Additionally, serum starvation increased the basal activity of the CLU promoter compared to cells grown in whole serum, suggesting a possible autocrine feedback loop induced by IR, where irradiated cells not only up- regulate IGF-1R, but also presumably increase production of the ligand, IGF-1.

Consistent with a previous report (384), we demonstrate induction of IGF-1R after IR

(Fig. 4D), as well as a 20% increase in secretion of IGF-1, 24-72 h post-IR (Fig. 4E).

Importantly, the induction of IGF-1R and its ligand provide a plausible explanation for the late induction of CLU after IR, as well as the increase in basal promoter activity after serum-starvation.

We show, for the first time, that the Src-Raf-Mek-Erk-1/2 pathway is required for

IR-induced sCLU activation. This cascade culminates in the activation of the Egr-1 transcription factor. In addition, we demonstrate a novel re-activation of the MAPK cascade after IR that correlates with the temporal activation of sCLU. The physiological relevance of this biphasic activation of MAPK is unknown. EGFR may be required for the initial induction of MAPK after IR, and inhibitors to EGFR are known to potentiate the cytotoxicity of radiation therapy (404). In addition, Lu et al., demonstrated that increased IGF-1R production in MCF-7 cells caused increased resistance to Herceptin (a monoclonal antibody to the Her2/neu receptor) induced cell death (405), suggesting a role for IGF-1R signaling in the development of resistance to this type of therapy. Our data suggests that tumor cells that survive the initial phase of treatment may develop resistance to EGFR inhibitors, as a result of the late induction of MAPK signaling through IGF-R1 upregulation, and potentially the upregulation of sCLU. This suggests

130 that the current antibody and small molecule therapies used to treat EGFR positive tumors may be optimized by the addition of inhibitors to the IGF-1R pathway.

Recent data indicate that sCLU provides cytoprotection against doxorubicin, taxol and cisplatin in the treatment of cancer cells (63, 73, 74, 87). In this report, we show that sCLU provides cytoprotection in MCF-7 cells after IR exposure (Fig. 1). sCLU expression has been found to be elevated in many types of tumors, including prostate, colorectal and breast cancer (49, 51, 58). In a recent paper by Chen et al., it was shown that CLU message and protein were elevated in intestinal tumors derived from mice containing the Apc min (multiple intestinal neoplasia) mutation (59). IGF-1R and IGF-1 production are also elevated in many tumor types. Gleave et al., in a recent review, suggest using CLU and insulin-like growth factor binding proteins (IGFBPs) as targets for antisense therapy against prostate cancer (73), although they did not mention a possible connection between IGF-1R signaling and the induction of sCLU. Our data strongly suggests such a connection, which may explain the connection between elevated sCLU levels and the development of prostate cancer.

It is intriguing to speculate a possible role for sCLU in bystander effects, after radiation or chemotherapeutic therapies. An increased production and secretion of sCLU by tumor cells into the lymph or vasculature system may provide a survival effect for neighboring or metastatic cancer cells. Additionally, secretion of IGF-1 by normal or tumor cells may provide a means for sCLU upregulation.

The signaling pathway that relieves p53 repression, allowing for sCLU induction after IR, has not yet been determined. Recently, in a paper by Lu et al., it was shown that

Src family kinases could inhibit the function of PTEN (412). PTEN has been shown to

131 be an inhibitor of IGF-1R activated MAPK (413) as well as the Akt signaling pathway

(412). One function of Akt is to stabilize Mdm2 allowing for degradation of p53.

Additionally, Tanno et al., have shown that Akt activation can up-regulate IGF-1R (414).

The cross-talk between the IGF-1R, PTEN and Akt pathways may provide an intriguing connection between the signaling cascade resulting in sCLU induction after IR and the repressive effects of p53 (Figure 4.1). Preliminary data from our laboratory support a role for PTEN and Akt in the regulation of sCLU. Over-expression of a constitutively active Akt or expression of a catalytically dead PTEN in 1403 cells resulted in significantly higher basal levels of CLU promoter activity as compared to cells transfected with vector alone (Criswell et al., unpublished data).

The exact regions of the CLU promoter that are required for IR induction have not yet been determined. We demonstrated that Egr-1 was required for sCLU induction, but there are three potential Egr-1 binding sites within the 1403 promoter fragment. Deletion analyses and point mutations of the CLU promoter will allow us to identify the sites required for IR-induction, as well as sites required for p53 repression. These experiments will also be useful in determining if the induction of sCLU by other cytotoxic agents occurs through the same signaling pathways as those induced by IR. Determining the exact element required for IR induction will allow us to potentially utilize the CLU promoter for combination gene targeting and radiation therapies. The Egr-1 promoter linked to the herpes simplex virus thymidine kinase gene is currently being utilized in this manner to sensitize tumor cells for radio-therapy (285, 415, 416). The CLU promoter IR- inducible element may be more useful since it is induced by much lower doses of IR and because it is regulated by p53. Since p53 is mutated in over 50% of all human tumors,

132 using the IR-inducible 4250 CLU promoter with the p53 binding site will allow us to target this construct to be active in the tumor cells missing p53, while sparing the normal cells that still contain wild-type p53.

This thesis has described the regulation of sCLU after IR. These studies shed light on the cellular responses to ionizing radiation. One of the problems facing cancer patients undergoing radio-therapy is the development of refractory cells/tumors to this treatment. sCLU may play a role in the development of radio-resistance, since cells lacking sCLU tend to be more sensitive to IR. Targeting the pathways that result in sCLU induction should sensitize tumor cells to radio-therapy. A greater understanding of these processes will allow for the improvement of cancer treatment.

133 IGF-1R

Src PI-3K IRS-1 SOS PTEN Grb2 Ras Raf Akt

MEK-1

Erk-2 Mdm2 Erk-1

Erk-2 Erk-1 p53 Egr-1 CLU promoter

Figure 4.1. Model depicting potential cross-talk between the IGF-1R dependent signaling pathway leading to sCLU induction after IR and a potential pathway signaling to the p53 transcriptional repression of sCLU.

134 BIBLIOGRAPHY:

1. Blaschuk O, Burdzy K, Fritz I 1983 Purification and characterization of a cell-

aggregating factor (clusterin), the major glycoprotein in ram rete testis fluid. J

Biol Chem 12:7714-7720

2. Fritz I, Burdzy K, Setchell B, Blaschuk O 1983 Ram rete testes fluid contains a

protein (clusterin) which influences cell-cell interactions in vitro. Biol Reprod

28:1173-1188

3. Leger JG, Montpetit ML, Tenniswood MP 1987 Characterization and cloning

of -repressed mRNAs from rat ventral prostate. Biochem Biophys Res

Commun 147:196-203

4. Bettuzzi S, Hiipakka RA, Gilna P, Liao ST 1989 Identification of an androgen-

repressed mRNA in rat ventral prostate as coding for sulphated glycoprotein 2 by

cDNA cloning and sequence analysis. Biochem J 257:293-6

5. de Silva HV, Stuart WD, Park YB, et al. 1990 Purification and characterization

of apolipoprotein J. J Biol Chem 265:14292-7

6. Fischer-Colbrie R, Zangerle R, Frischenschlager I, Weber A, Winkler H

1984 Isolation and immunological characterization of a glycoprotein from adrenal

chromaffin granules. J Neurochem 42:1008-16

7. Hartmann K, Rauch J, Urban J, et al. 1991 Molecular cloning of gp 80, a

glycoprotein complex secreted by kidney cells in vitro and in vivo. A link to the

reproductive system and to the complement cascade. J Biol Chem 266:9924-31

135 8. James RW, Hochstrasser AC, Borghini I, Martin B, Pometta D, Hochstrasser

D 1991 Characterization of a human high density lipoprotein-associated protein,

NA1/NA2. Identity with SP-40,40, an inhibitor of complement-mediated

cytolysis. Arterioscler Thromb 11:645-52

9. Jenne DE, Lowin B, Peitsch MC, Bottcher A, Schmitz G, Tschopp J 1991

Clusterin (complement lysis inhibitor) forms a high density lipoprotein complex

with apolipoprotein A-I in human plasma. J Biol Chem 266:11030-6.

10. Kirszbaum L, Sharpe JA, Murphy B, et al. 1989 Molecular cloning and

characterization of the novel, human complement-associated protein, SP-40,40: a

link between the complement and reproductive systems. Embo J 8:711-8

11. Boothman DA, Bouvard I, Hughes EN 1989 Identification and characterization

of X-ray-induced proteins in human cells. Cancer Res 49:2871-8

12. Burkey BF, Stuart WD, Harmony JA 1992 Hepatic apolipoprotein J is secreted

as a lipoprotein. J Lipid Res 33:1517-26

13. Calero M, Tokuda T, Rostagno A, et al. 1999 Functional and structural

properties of lipid-associated apolipoprotein J (clusterin). Biochem J 344 Pt

2:375-83

14. Murphy BF, Saunders JR, O'Bryan MK, Kirszbaum L, Walker ID, d'Apice

AJ 1989 SP-40,40 is an inhibitor of C5b-6-initiated haemolysis. Int Immunol

1:551-4

15. Tschopp J, Chonn A, Hertig S, French LE 1993 Clusterin, the human

apolipoprotein and complement inhibitor, binds to complement C7, C8 beta, and

the b domain of C9. J Immunol 151:2159-65

136 16. Moulson CL, Millis AJ 1999 Clusterin (Apo J) regulates vascular smooth muscle

cell differentiation in vitro. J Cell Physiol 180:355-64

17. Ahuja HS, Tenniswood M, Lockshin R, Zakeri ZF 1994 Expression of

clusterin in cell differentiation and cell death. Biochem Cell Biol 72:523-30

18. O'Bryan MK, Baker HW, Saunders JR, et al. 1990 Human seminal clusterin

(SP-40,40). Isolation and characterization. J Clin Invest 85:1477-86

19. de Silva HV, Harmony JA, Stuart WD, Gil CM, Robbins J 1990

Apolipoprotein J: structure and tissue distribution. Biochemistry 29:5380-9

20. Michel D, Chabot JG, Moyse E, Danik M, Quirion R 1992 Possible functions

of a new genetic marker in central nervous system: the sulfated glycoprotein-2

(SGP-2). Synapse 11:105-11

21. Guenette RS, Corbeil HB, Leger J, et al. 1994 Induction of gene expression

during involution of the lactating mammary gland of the rat. J Mol Endocrinol

12:47-60

22. Bonnard AS, Chan P, Fontaine M 1997 Expression of clusterin and C4 mRNA

during rat peripheral nerve regeneration. Immunopharmacology 38:81-6

23. Bursch W, Gleeson T, Kleine L, Tenniswood M 1995 Expression of clusterin

(testosterone-repressed prostate message-2) mRNA during growth and

regeneration of rat liver. Arch Toxicol 69:253-8

24. Gobe GC, Buttyan R, Wyburn KR, Etheridge MR, Smith PJ 1995 Clusterin

expression and apoptosis in tissue remodeling associated with renal regeneration.

Kidney Int 47:411-20

137 25. Helvering LM, Richardson KK, Horn DM, et al. 1993 Expression of TRPM-2

during involution and regeneration of the rat liver. Cancer Lett 71:133-42

26. Fratelli M, Galli G, Minto M, Pasinetti GM 1996 Role of clusterin in cell

adhesion during early phases of in P19 embryonic

carcinoma cells. Biochim Biophys Acta 1311:71-6

27. Silkensen JR, Skubitz AP, Skubitz KM, Rosenberg ME 1999 Identification of

clusterin sequences mediating renal tubular cell interactions. J Pept Res 54:449-57

28. Silkensen JR, Skubitz KM, Skubitz AP, et al. 1995 Clusterin promotes the

aggregation and adhesion of renal porcine epithelial cells. J Clin Invest 96:2646-

53

29. Buttyan R, Olsson CA, Pintar J, et al. 1989 Induction of the TRPM-2 gene in

cells undergoing programmed death. Mol Cell Biol 9:3473-81

30. Oliveira AS, Corbo M, Duigou G, Gabbai AA, Hays AP 1993 Expression of a

cell death marker (Clusterin) in muscle target fibers. Arq Neuropsiquiatr 51:371-6

31. Tenniswood MP, Guenette RS, Lakins J, Mooibroek M, Wong P, Welsh JE

1992 Active cell death in hormone-dependent tissues. Cancer Rev

11:197-220

32. Pearse MJ, O'Bryan M, Fisicaro N, Rogers L, Murphy B, d'Apice AJ 1992

Differential expression of clusterin in inducible models of apoptosis. Int Immunol

4:1225-31

33. Dragunow M, Preston K, Dodd J, Young D, Lawlor P, Christie D 1995

Clusterin accumulates in dying neurons following status epilepticus. Brain Res

Mol Brain Res 32:279-90

138 34. Lakins J, Bennett SA, Chen JH, et al. 1998 Clusterin biogenesis is altered

during apoptosis in the regressing rat ventral prostate. J Biol Chem 273:27887-95

35. Yang CR, Leskov K, Hosley-Eberlein K, et al. 2000 Nuclear clusterin/XIP8, an

x-ray-induced Ku70-binding protein that signals cell death. Proc Natl Acad Sci U

S A 97:5907-12

36. Jordan-Starck TC, Lund SD, Witte DP, et al. 1994 Mouse apolipoprotein J:

characterization of a gene implicated in atherosclerosis. J Lipid Res 35:194-210

37. Gelissen IC, Hochgrebe T, Wilson MR, et al. 1998 Apolipoprotein J (clusterin)

induces cholesterol export from macrophage-foam cells: a potential anti-

atherogenic function? Biochem J 331:231-7

38. Silkensen JR, Schwochau GB, Rosenberg ME 1994 The role of clusterin in

tissue injury. Biochem Cell Biol 72:483-8

39. Calero M, Rostagno A, Matsubara E, Zlokovic B, Frangione B, Ghiso J 2000

Apolipoprotein J (clusterin) and Alzheimer's disease. Microsc Res Tech 50:305-

315

40. Zlokovic BV 1996 Cerebrovascular transport of Alzheimer's amyloid beta and

J and E: possible anti-amyloidogenic role of the blood-brain

barrier. Life Sci 59:1483-97

41. Lidstrom AM, Bogdanovic N, Hesse C, Volkman I, Davidsson P, Blennow K

1998 Clusterin (apolipoprotein J) protein levels are increased in hippocampus and

in frontal cortex in Alzheimer's disease. Exp Neurol 154:511-21.

42. Dvergsten J, Manivel JC, Correa-Rotter R, Rosenberg ME 1994 Expression

of clusterin in human renal diseases. Kidney Int 45:828-35

139 43. Nath KA, Dvergsten J, Correa-Rotter R, Hostetter TH, Manivel JC,

Rosenberg ME 1994 Induction of clusterin in acute and chronic oxidative renal

disease in the rat and its dissociation from cell injury. Lab Invest 71:209-18

44. Watanabe H, Hamada H, Yamada N, et al. 2004 Proteome analysis reveals

elevated serum levels of clusterin in patients with preeclampsia. Proteomics

4:537-43

45. Newkirk MM, Apostolakos P, Neville C, Fortin PR 1999 Systemic lupus

erythematosus, a disease associated with low levels of clusterin/apoJ, an

antiinflammatory protein. J Rheumatol 26:597-603

46. Jones SE, Meerabux JM, Yeats DA, Neal MJ 1992 Analysis of differentially

expressed genes in retinitis pigmentosa retinas. Altered expression of clusterin

mRNA. FEBS Lett 300:279-82

47. Wong P, Borst DE, Farber D, et al. 1994 Increased TRPM-2/clusterin mRNA

levels during the time of retinal degeneration in mouse models of retinitis

pigmentosa. Biochem Cell Biol 72:439-46

48. Miyake H, Hara I, Kamidono S, Gleave ME, Eto H 2003 Resistance to

cytotoxic chemotherapy-induced apoptosis in human prostate cancer cells is

associated with intracellular clusterin expression. Oncol Rep 10:469-73

49. Steinberg J, Oyasu R, Lang S, et al. 1997 Intracellular levels of SGP-2

(Clusterin) correlate with tumor grade in prostate cancer. Clin Cancer Res 3:1707-

11

50. Tenniswood M 1997 Apoptosis, tumour invasion and prostate cancer. Br J Urol

79 Suppl 2:27-34

140 51. Xie MJ, Motoo Y, Su SB, et al. 2002 Expression of clusterin in human

pancreatic cancer. Pancreas 25:234-8

52. Hasan Q, Ruger BM, Tan ST, Gush J, Davis PF 2000 Clusterin/apoJ

expression during the development of hemangioma. Hum Pathol 31:691-7

53. Saffer H, Wahed A, Rassidakis GZ, Medeiros LJ 2002 Clusterin expression in

malignant lymphomas: a survey of 266 cases. Mod Pathol 15:1221-6

54. Lae ME, Ahmed I, Macon WR 2002 Clusterin is widely expressed in systemic

anaplastic large cell lymphoma but fails to differentiate primary from secondary

cutaneous anaplastic large cell lymphoma. Am J Clin Pathol 118:773-9

55. Wellmann A, Thieblemont C, Pittaluga S, et al. 2000 Detection of

differentially expressed genes in lymphomas using cDNA arrays: identification of

clusterin as a new diagnostic marker for anaplastic large-cell lymphomas. Blood

96:398-404

56. Zwain I, Amato P 2000 Clusterin protects granulosa cells from apoptotic cell

death during follicular atresia. Exp Cell Res 257:101-10

57. Hough CD, Cho KR, Zonderman AB, Schwartz DR, Morin PJ 2001

Coordinately up-regulated genes in ovarian cancer. Cancer Res 61:3869-76.

58. Redondo M, Villar E, Torres-Munoz J, Tellez T, Morell M, Petito CK 2000

Overexpression of clusterin in human breast carcinoma. Am J Pathol 157:393-9

59. Chen X, Halberg RB, Ehrhardt WM, Torrealba J, Dove WF 2003 Clusterin as

a biomarker in murine and human intestinal neoplasia. Proc Natl Acad Sci U S A

100:9530-5

141 60. Hara I, Miyake H, Gleave ME, Kamidono S 2001 Introduction of Clusterin

Gene into Human Renal Cell Carcinoma Cells Enhances Their Resistance to

Cytotoxic Chemotherapy through Inhibition of Apoptosis both in vitro and in

vivo. Jpn J Cancer Res 92:1220-4.

61. Zellweger T, Miyake H, July LV, Akbari M, Kiyama S, Gleave ME 2001

Chemosensitization of human renal cell cancer using antisense oligonucleotides

targeting the antiapoptotic gene clusterin. Neoplasia 3:360-7.

62. Miyake H, Chi KN, Gleave ME 2000 Antisense TRPM-2 oligodeoxynucleotides

chemosensitize human androgen-independent PC-3 prostate cancer cells both in

vitro and in vivo. Clin Cancer Res 6:1655-63

63. Chung J, Kwak C, Jin RJ, Lee CH, Lee KH, Lee SE 2004 Enhanced

chemosensitivity of bladder cancer cells to cisplatin by suppression of clusterin in

vitro. Cancer Lett 203:155-61

64. Bandyk MG, Sawczuk IS, Olsson CA, Katz AE, Buttyan R 1990

Characterization of the products of a gene expressed during androgen-

programmed cell death and their potential use as a marker of urogenital injury. J

Urol 143:407-13

65. Danik M, Chabot JG, Mercier C, et al. 1991 Human gliomas and epileptic foci

express high levels of a mRNA related to rat testicular sulfated glycoprotein 2, a

purported marker of cell death. Proc Natl Acad Sci U S A 88:8577-81

66. Michel D, Gillet G, Volovitch M, Pessac B, Calothy G, Brun G 1989

Expression of a novel gene encoding a 51.5 kD precursor protein is induced by

different retroviral oncogenes in quail neuroretinal cells. Oncogene Res 4:127-36

142 67. Boggs LN, Fuson KS, Baez M, et al. 1996 Clusterin (Apo J) protects against in

vitro amyloid-beta (1-40) neurotoxicity. J Neurochem 67:1324-7

68. de Silva HV, Stuart WD, Duvic CR, et al. 1990 A 70-kDa apolipoprotein

designated ApoJ is a marker for subclasses of human plasma high density

lipoproteins. J Biol Chem 265:13240-7

69. Herault Y, Chatelain G, Brun G, Michel D 1992 V-src-induced-transcription of

the avian clusterin gene. Nucleic Acids Res 20:6377-83

70. Diemer V, Hoyle M, Baglioni C, Millis AJ 1992 Expression of porcine

complement cytolysis inhibitor mRNA in cultured aortic smooth muscle cells.

Changes during differentiation in vitro. J Biol Chem 267:5257-64

71. Yang CR, Yeh S, Leskov K, et al. 1999 Isolation of Ku70-binding proteins

(KUBs). Nucleic Acids Res 27:2165-74

72. Miyake H, Hara I, Kamidono S, Gleave ME 2001 Synergistic

chemsensitization and inhibition of tumor growth and metastasis by the antisense

oligodeoxynucleotide targeting clusterin gene in a human bladder cancer model.

Clin Cancer Res 7:4245-52.

73. Gleave ME, Miyake H, Zellweger T, et al. 2001 Use of antisense

oligonucleotides targeting the antiapoptotic gene, clusterin/testosterone-repressed

prostate message 2, to enhance androgen sensitivity and chemosensitivity in

prostate cancer. Urology 58:39-49.

74. Zellweger T, Miyake H, Cooper S, et al. 2001 Antitumor activity of antisense

clusterin oligonucleotides is improved in vitro and in vivo by incorporation of 2'-

O-(2-methoxy)ethyl chemistry. J Pharmacol Exp Ther 298:934-40.

143 75. Humphreys DT, Carver JA, Easterbrook-Smith SB, Wilson MR 1999

Clusterin has chaperone-like activity similar to that of small heat shock proteins. J

Biol Chem 274:6875-81

76. Wilson MR, Easterbrook-Smith SB 2000 Clusterin is a secreted mammalian

chaperone. Trends Biochem Sci 25:95-8

77. Poon S, Easterbrook-Smith SB, Rybchyn MS, Carver JA, Wilson MR 2000

Clusterin is an ATP-independent chaperone with very broad substrate specificity

that stabilizes stressed proteins in a folding-competent state. Biochemistry

39:15953-60.

78. Poon S, Treweek TM, Wilson MR, Easterbrook-Smith SB, Carver JA 2002

Clusterin is an extracellular chaperone that specifically interacts with slowly

aggregating proteins on their off-folding pathway. FEBS Lett 513:259-66

79. Poon S, Rybchyn MS, Easterbrook-Smith SB, Carver JA, Pankhurst GJ,

Wilson MR 2002 Mildly acidic pH activates the extracellular molecular

chaperone clusterin. J Biol Chem 277:39532-40

80. McLaughlin L, Zhu G, Mistry M, et al. 2000 Apolipoprotein J/clusterin limits

the severity of murine autoimmune myocarditis. J Clin Invest 106:1105-13.

81. Han BH, DeMattos RB, Dugan LL, et al. 2001 Clusterin contributes to caspase-

3-independent brain injury following neonatal hypoxia-ischemia. Nat Med 7:338-

43.

82. Rosenberg ME, Girton R, Finkel D, et al. 2002 Apolipoprotein J/clusterin

prevents a progressive glomerulopathy of aging. Mol Cell Biol 22:1893-902

144 83. DeMattos RB, Cirrito JR, Parsadanian M, et al. 2004 ApoE and Clusterin

Cooperatively Suppress Abeta Levels and Deposition. Evidence that ApoE

Regulates Extracellular Abeta Metabolism In Vivo. Neuron 41:193-202

84. Jin G, Howe PH 1997 Regulation of clusterin gene expression by transforming

growth factor beta. J Biol Chem 272:26620-6

85. Wegrowski Y, Perreau C, Martiny L, Haye B, Maquart FX, Bellon G 1999

Transforming growth factor beta-1 up-regulates clusterin synthesis in thyroid

epithelial cells. Exp Cell Res 247:475-83

86. Jin G, Howe PH 1999 Transforming growth factor beta regulates clusterin gene

expression via modulation of transcription factor c-Fos. Eur J Biochem 263:534-

42

87. Cervellera M, Raschella G, Santilli G, et al. 2000 Direct transactivation of the

anti-apoptotic gene apolipoprotein J (clusterin) by B-MYB. J Biol Chem

275:21055-60

88. Saura J, Petegnief V, Wu X, Liang Y, Paul SM 2003 Microglial apolipoprotein

E and astroglial apolipoprotein J expression in vitro: opposite effects of

lipopolysaccharide. J Neurochem 85:1455-67

89. Li X, Massa PE, Hanidu A, et al. 2002 IKKalpha, IKKbeta, and

NEMO/IKKgamma are each required for the NF-kappa B-mediated inflammatory

response program. J Biol Chem 277:45129-40

90. Santilli G, Aronow BJ, Sala A 2003 Essential requirement of apolipoprotein J

(clusterin) signaling for IkappaB expression and regulation of NF-kappaB

activity. J Biol Chem 278:38214-9

145 91. Boothman DA, Meyers M, Fukunaga N, Lee SW 1993 Isolation of x-ray-

inducible transcripts from radioresistant human melanoma cells. Proc Natl Acad

Sci U S A 90:7200-4

92. Boothman DA, Majmudar G, Johnson T 1994 Immediate X-ray-inducible

responses from mammalian cells. Radiat Res 138:S44-6

93. Wong P, Pineault J, Lakins J, et al. 1993 Genomic organization and expression

of the rat TRPM-2 (clusterin) gene, a gene implicated in apoptosis. J Biol Chem

268:5021-31

94. Criswell T, Klokov D, Beman M, Lavik JP, Boothman DA 2003 Repression of

IR-inducible clusterin expression by the p53 tumor suppressor protein. Cancer

Biol Ther 2:372-80

95. Leskov KS, Klokov DY, Li J, Kinsella TJ, Boothman DA 2003 Synthesis and

functional analyses of nuclear clusterin, a cell death protein. J Biol Chem

278:11590-600

96. Wallace SS 2002 Biological consequences of free radical-damaged DNA bases.

Free Radic Biol Med 33:1-14

97. Davis TW, Wilson-Van Patten C, Yang C-R, et al. 1998 Transcriptional

responses to damage created by ionizing radiation. In: Nickoloff JAaH, M. (ed)

DNA Damage and Repair. Humana Press, Inc., Totowa, NJ, pp 223-261

98. Jackson SP 1996 DNA damage detection by DNA dependent protein kinase and

related enzymes. Cancer Surv 28:261-79

99. Jackson SP 1997 DNA-dependent protein kinase. Int J Biochem Cell Biol

29:935-8

146 100. Barlow C, Brown KD, Deng CX, Tagle DA, Wynshaw-Boris A 1997 Atm

selectively regulates distinct p53-dependent cell-cycle checkpoint and apoptotic

pathways. Nat Genet 17:453-6

101. Shiloh Y 1997 Ataxia-telangiectasia and the Nijmegen breakage syndrome:

related disorders but genes apart. Annu Rev Genet 31:635-62

102. Tibbetts RS, Brumbaugh KM, Williams JM, et al. 1999 A role for ATR in the

DNA damage-induced phosphorylation of p53. Genes Dev 13:152-7

103. Tibbetts RS, Cortez D, Brumbaugh KM, et al. 2000 Functional interactions

between BRCA1 and the checkpoint kinase ATR during genotoxic stress. Genes

Dev 14:2989-3002

104. Carney JP, Maser RS, Olivares H, et al. 1998 The hMre11/hRad50 protein

complex and Nijmegen breakage syndrome: linkage of double-strand break repair

to the cellular DNA damage response. Cell 93:477-86

105. Fishel R 1998 Mismatch repair, molecular switches, and signal transduction.

Genes Dev 12:2096-101

106. Fishel R 2001 The selection for mismatch repair defects in hereditary

nonpolyposis colorectal cancer: revising the mutator hypothesis. Cancer Res

61:7369-74

107. Kolodner R 1996 Biochemistry and genetics of eukaryotic mismatch repair.

Genes Dev 10:1433-42

108. Lindsey-Boltz LA, Bermudez VP, Hurwitz J, Sancar A 2001 Purification and

characterization of human DNA damage checkpoint Rad complexes. Proc Natl

Acad Sci U S A 98:11236-41

147 109. Burtelow MA, Roos-Mattjus PM, Rauen M, Babendure JR, Karnitz LM

2001 Reconstitution and molecular analysis of the hRad9-hHus1-hRad1 (9-1-1)

DNA damage responsive checkpoint complex. J Biol Chem 276:25903-9

110. Rauen M, Burtelow MA, Dufault VM, Karnitz LM 2000 The human

checkpoint protein hRad17 interacts with the PCNA-like proteins hRad1, hHus1,

and hRad9. J Biol Chem 275:29767-71

111. Kasid U, Wang FH, Whiteside TL 1997 Ionizing radiation and TNF-alpha

stimulate gene expression of a Thr/Tyr-protein phosphatase HVH1 and inhibitory

factor IkappaB alpha in human squamous carcinoma cells. Mol Cell Biochem

173:193-7

112. Fornace AJ, Jr., Nebert DW, Hollander MC, et al. 1989 Mammalian genes

coordinately regulated by growth arrest signals and DNA-damaging agents. Mol

Cell Biol 9:4196-203

113. Appella E 2001 Modulation of p53 function in cellular regulation. Eur J Biochem

268:2763

114. Appella E, Anderson CW 2001 Post-translational modifications and activation

of p53 by genotoxic stresses. Eur J Biochem 268:2764-72

115. Stewart ZA, Pietenpol JA 2001 p53 Signaling and cell cycle checkpoints. Chem

Res Toxicol 14:243-63

116. Vogelstein B, Lane D, Levine AJ 2000 Surfing the p53 network. Nature

408:307-10

117. Prives C, Hall PA 1999 The p53 pathway. J Pathol 187:112-26

148 118. Khanna KK, Lavin MF 1993 Ionizing radiation and UV induction of p53

protein by different pathways in ataxia-telangiectasia cells. Oncogene 8:3307-12

119. Graeber TG, Peterson JF, Tsai M, Monica K, Fornace AJ, Jr., Giaccia AJ

1994 Hypoxia induces accumulation of p53 protein, but activation of a G1-phase

checkpoint by low-oxygen conditions is independent of p53 status. Mol Cell Biol

14:6264-77

120. Zhan Q, Carrier F, Fornace AJ, Jr. 1993 Induction of cellular p53 activity by

DNA-damaging agents and growth arrest. Mol Cell Biol 13:4242-50

121. Abraham RT 2001 Cell cycle checkpoint signaling through the ATM and ATR

kinases. Genes Dev 15:2177-96

122. Wahl GM, Carr AM 2001 The evolution of diverse biological responses to DNA

damage: insights from yeast and p53. Nat Cell Biol 3:E277-86

123. Pettitt AR, Sherrington PD, Stewart G, Cawley JC, Taylor AM, Stankovic T

2001 p53 dysfunction in B-cell chronic lymphocytic leukemia: inactivation of

ATM as an alternative to TP53 mutation. Blood 98:814-22

124. Lu X, Lane DP 1993 Differential induction of transcriptionally active p53

following UV or ionizing radiation: defects in instability

syndromes? Cell 75:765-78.

125. Zhan Q, Bae I, Kastan MB, Fornace AJ, Jr. 1994 The p53-dependent gamma-

ray response of GADD45. Cancer Res 54:2755-60

126. Brach MA, Gruss HJ, Kaisho T, Asano Y, Hirano T, Herrmann F 1993

Ionizing radiation induces expression of interleukin 6 by human fibroblasts

involving activation of nuclear factor-kappa B. J Biol Chem 268:8466-72

149 127. Sahijdak WM, Yang CR, Zuckerman JS, Meyers M, Boothman DA 1994

Alterations in transcription factor binding in radioresistant human melanoma cells

after ionizing radiation. Radiat Res 138:S47-51

128. Yang CR, Wilson-Van Patten C, Planchon SM, et al. 2000 Coordinate

modulation of Sp1, NF-kappa B, and p53 in confluent human malignant

melanoma cells after ionizing radiation. Faseb J 14:379-90

129. Datta R, Taneja N, Sukhatme VP, Qureshi SA, Weichselbaum R, Kufe DW

1993 Reactive oxygen intermediates target CC(A/T)6GG sequences to mediate

activation of the early growth response 1 transcription factor gene by ionizing

radiation. Proc Natl Acad Sci U S A 90:2419-22

130. Hallahan DE, Gius D, Kuchibhotla J, Sukhatme V, Kufe DW, Weichselbaum

RR 1993 Radiation signaling mediated by Jun activation following dissociation

from a cell type-specific repressor. J Biol Chem 268:4903-7

131. Sherman ML, Datta R, Hallahan DE, Weichselbaum RR, Kufe DW 1990

Ionizing radiation regulates expression of the c-jun protooncogene. Proc Natl

Acad Sci U S A 87:5663-6

132. Meighan-Mantha RL, Riegel AT, Suy S, et al. 1999 Ionizing radiation

stimulates octamer factor DNA binding activity in human carcinoma cells. Mol

Cell Biochem 199:209-15

133. Unger T, Mietz JA, Scheffner M, Yee CL, Howley PM 1993 Functional

domains of wild-type and mutant p53 proteins involved in transcriptional

regulation, transdominant inhibition, and transformation suppression. Mol Cell

Biol 13:5186-94

150 134. Wang P, Reed M, Wang Y, et al. 1994 p53 domains: structure, oligomerization,

and transformation. Mol Cell Biol 14:5182-91

135. Oliner JD, Pietenpol JA, Thiagalingam S, Gyuris J, Kinzler KW, Vogelstein

B 1993 Oncoprotein MDM2 conceals the activation domain of tumour suppressor

p53. Nature 362:857-60

136. Stewart GS, Last JI, Stankovic T, et al. 2001 Residual ataxia telangiectasia

mutated protein function in cells from ataxia telangiectasia patients, with

5762ins137 and 7271T-->G mutations, showing a less severe phenotype. J Biol

Chem 276:30133-41.

137. Bargonetti J, Manfredi JJ, Chen X, Marshak DR, Prives C 1993 A proteolytic

fragment from the central region of p53 has marked sequence-specific DNA-

binding activity when generated from wild-type but not from oncogenic mutant

p53 protein. Genes Dev 7:2565-74

138. el-Deiry WS, Kern SE, Pietenpol JA, Kinzler KW, Vogelstein B 1992

Definition of a consensus binding site for p53. Nat Genet 1:45-9

139. Pavletich NP, Chambers KA, Pabo CO 1993 The DNA-binding domain of p53

contains the four conserved regions and the major mutation hot spots. Genes Dev

7:2556-64

140. Fang S, Jensen JP, Ludwig RL, Vousden KH, Weissman AM 2000 Mdm2 is a

RING finger-dependent ubiquitin protein ligase for itself and p53. J Biol Chem

275:8945-51

141. Honda R, Tanaka H, Yasuda H 1997 Oncoprotein MDM2 is a

E3 for tumor suppressor p53. FEBS Lett 420:25-7

151 142. Michael D, Oren M 2003 The p53-Mdm2 module and the ubiquitin system.

Semin Cancer Biol 13:49-58

143. Mayo LD, Turchi JJ, Berberich SJ 1997 Mdm-2 phosphorylation by DNA-

dependent protein kinase prevents interaction with p53. Cancer Res 57:5013-6

144. Moll UM, Riou G, Levine AJ 1992 Two distinct mechanisms alter p53 in breast

cancer: mutation and nuclear exclusion. Proc Natl Acad Sci U S A 89:7262-6

145. Moll UM, LaQuaglia M, Benard J, Riou G 1995 Wild-type p53 protein

undergoes cytoplasmic sequestration in undifferentiated neuroblastomas but not in

differentiated tumors. Proc Natl Acad Sci U S A 92:4407-11

146. Roth J, Dobbelstein M, Freedman DA, Shenk T, Levine AJ 1998 Nucleo-

cytoplasmic shuttling of the hdm2 oncoprotein regulates the levels of the p53

protein via a pathway used by the human immunodeficiency virus rev protein.

Embo J 17:554-64

147. Hupp TR, Meek DW, Midgley CA, Lane DP 1992 Regulation of the specific

DNA binding function of p53. Cell 71:875-86

148. Waterman MJ, Stavridi ES, Waterman JL, Halazonetis TD 1998 ATM-

dependent activation of p53 involves dephosphorylation and association with 14-

3-3 proteins. Nat Genet 19:175-8

149. Stavridi ES, Chehab NH, Malikzay A, Halazonetis TD 2001 Substitutions that

compromise the ionizing radiation-induced association of p53 with 14-3-3

proteins also compromise the ability of p53 to induce cell cycle arrest. Cancer Res

61:7030-3

152 150. Saito S, Goodarzi AA, Higashimoto Y, et al. 2002 ATM mediates

phosphorylation at multiple p53 sites, including Ser(46), in response to ionizing

radiation. J Biol Chem 277:12491-4

151. Muller S, Berger M, Lehembre F, Seeler JS, Haupt Y, Dejean A 2000 c-Jun

and p53 activity is modulated by SUMO-1 modification. J Biol Chem 275:13321-

9

152. Lees-Miller SP, Sakaguchi K, Ullrich SJ, Appella E, Anderson CW 1992

Human DNA-activated protein kinase phosphorylates serines 15 and 37 in the

amino-terminal transactivation domain of human p53. Mol Cell Biol 12:5041-9

153. Araki R, Fukumura R, Fujimori A, et al. 1999 Enhanced phosphorylation of

p53 serine 18 following DNA damage in DNA-dependent protein kinase catalytic

subunit-deficient cells. Cancer Res 59:3543-6

154. Abraham J, Spaner D, Benchimol S 1999 Phosphorylation of p53 protein in

response to ionizing radiation occurs at multiple sites in both normal and DNA-

PK deficient cells. Oncogene 18:1521-7.

155. Jimenez GS, Bryntesson F, Torres-Arzayus MI, et al. 1999 DNA-dependent

protein kinase is not required for the p53-dependent response to DNA damage.

Nature 400:81-3

156. Banin S, Moyal L, Shieh S, et al. 1998 Enhanced phosphorylation of p53 by

ATM in response to DNA damage. Science 281:1674-7

157. Canman CE, Kastan MB 1998 Small contribution of G1 checkpoint control

manipulation to modulation of p53-mediated apoptosis. Oncogene 16:957-66

153 158. Hirao A, Kong YY, Matsuoka S, et al. 2000 DNA damage-induced activation of

p53 by the checkpoint kinase Chk2. Science 287:1824-7

159. Ahn JY, Schwarz JK, Piwnica-Worms H, Canman CE 2000 Threonine 68

phosphorylation by ataxia telangiectasia mutated is required for efficient

activation of Chk2 in response to ionizing radiation. Cancer Res 60:5934-6

160. Fan F, Jin S, Amundson SA, et al. 2002 ATF3 induction following DNA

damage is regulated by distinct signaling pathways and over-expression of ATF3

protein suppresses cells growth. Oncogene 21:7488-96

161. Embree-Ku M, Venturini D, Boekelheide K 2002 Fas is involved in the p53-

dependent apoptotic response to ionizing radiation in mouse testis. Biol Reprod

66:1456-61

162. Burger H, Nooter K, Boersma AW, Kortland CJ, van den Berg AP, Stoter G

1998 Expression of p53, p21/WAF/CIP, Bcl-2, Bax, Bcl-x, and Bak in radiation-

induced apoptosis in testicular germ cell tumor lines. Int J Radiat Oncol Biol Phys

41:415-24

163. Fei P, Bernhard EJ, El-Deiry WS 2002 Tissue-specific induction of p53 targets

in vivo. Cancer Res 62:7316-27

164. Lin Y, Ma W, Benchimol S 2000 Pidd, a new death-domain-containing protein,

is induced by p53 and promotes apoptosis. Nat Genet 26:122-7

165. Takai H, Naka K, Okada Y, et al. 2002 Chk2-deficient mice exhibit

radioresistance and defective p53-mediated transcription. Embo J 21:5195-205

166. Sen R, Baltimore D 1986 Multiple nuclear factors interact with the

immunoglobulin enhancer sequences. Cell 46:705-16

154 167. Miyamoto S, Verma IM 1995 Rel/NF-kB/IkB story. Adv. Cancer Res. 66:255-

92

168. Verma IM, Stevenson JK, Schwarz EM, Van Antwerp D, Miyamoto S 1995

Rel/NF-kB/IkB family: intimate tales of association and dissociation. Genes Dev.

9:2723-35

169. Karin M, Ben-Neriah Y 2000 Phosphorylation meets ubiquitination: the control

of NF-kB activity. Annu. Rev. Immunol. 18:621-63

170. Silverman N, Maniatis T 2001 NF-kB signaling pathways in mammalian and

insect innate immunity. Genes Dev. 15:2321-42

171. Ghosh S, May MJ, Kopp EB 1998 NF-kB and Rel proteins: evolutionarily

conserved mediators of immune responses. Annu. Rev. Immunol. 16:225-60

172. Ghosh S, Karin M 2002 Missing pieces in the NF-kB puzzle. Cell 109:S81-96

173. Prasad AV, Mohan N, Chandrasekar B, Meltz ML 1994 Activation of nuclear

factor kappa B in human lymphoblastoid cells by low-dose ionizing radiation.

Radiat Res 138:367-72

174. Ashburner BP, Shackelford RE, Baldwin AS, Jr., Paules RS 1999 Lack of

involvement of ataxia telangiectasia mutated (ATM) in regulation of nuclear

factor-kappaB (NF-kappaB) in human diploid fibroblasts. Cancer Res 59:5456-60

175. Zhou D, Brown SA, Yu T, et al. 1999 A high dose of ionizing radiation induces

tissue-specific activation of nuclear factor-kappaB in vivo. Rad. Res. 151:703-9

176. Li N, Karin M 1999 Is NF-kB the sensor of oxidative stress? FASEB J. 13:1137-

43

155 177. Raju U, Gumin GJ, Tofilon PJ 2000 Radiation-induced transcription factor

activation in the rat cerebral cortex. Int. J. Radiat. Biol. 76:1045-53

178. Huang TT, Wuerzberger-Davis SM, Seufzer BJ, et al. 2000 NF-kB activation

by camptothecin. A linkage between nuclear DNA damage and cytoplasmic

signaling events. J. Biol. Chem. 275:9501-9

179. Zhou BB, Elledge SJ 2000 The DNA damage response: putting checkpoints in

perspective. Nature 408:433-9

180. Li N, Banin S, Ouyang H, et al. 2001 ATM is required for IkB kinase (IKK)

activation in response to DNA double strand breaks. J. Biol. Chem. 276:8898-903

181. Jung M, Zhang Y, Lee S, Dritschilo A 1995 Correction of radiation sensitivity

in ataxia telangiectasia cells by a truncated IkBa. Science 268:1619-21

182. Jung M, Kondratyev A, Lee SA, Dimtchev A, Dritschilo A 1997 ATM gene

product phosphorylates IkBa. Cancer Res. 57:24-7

183. Lee SA, Dritschilo A, Jung M 1998 Impaired ionizing radiation-induced

activation of a nuclear signal essential for phosphorylation of c-Jun by dually

phosphorylated c-Jun amino-terminal kinases in ataxia telangiectasia fibroblasts. J

Biol Chem 273:32889-94.

184. Basu S, Rosenzweig KR, Youmell M, Price BD 1998 The DNA-dependent

protein kinase participates in the activation of NF kB following DNA damage.

Biochem. Biophys. Res. Commun. 247:79-83

185. Li Z, Li JJ 2001 Effector genes altered in mcf-7 human breast cancer cells after

exposure to fractionated ionizing radiation. Radiat Res 155:543-53

156 186. Curry HA, Clemens RA, Shah S, et al. 1999 Heat shock inhibits radiation-

induced activation of NF-kappaB via inhibition of I-kappaB kinase. J Biol Chem

274:23061-7

187. Raju U, Gumin GJ, Noel F, Tofilon PJ 1998 IkBa degradation is not a

requirement for the X-ray-induced activation of nuclear factor kappaB in normal

rat astrocytes and human brain tumour cells. Int. J. Radiat. Biol. 74:617-24

188. Miyakoshi J, Yagi K 2000 Inhibition of IkBa phosphorylation at serine and

tyrosine acts independently on sensitization to DNA damaging agents in human

glioma cells. Br. J. Cancer 82:28-33

189. Shao R, Tsai EM, Wei K, et al. 2001 E1A inhibition of radiation-induced NF-kB

activity through suppression of IKK activity and IkB degradation, independent of

Akt activation. Cancer Res. 61:7413-6

190. Russell JS, Tofilon PJ 2002 Radiation-induced activation of nuclear factor-kB

involves selective degradation of plasma membrane-associated IkBa. Mol. Biol.

Cell. 13:3431-40

191. Uckun FM, Schieven GL, Tuel-Ahlgren LM, et al. 1993 Tyrosine

phosphorylation is a mandatory proximal step in radiation-induced activation of

the protein kinase C signaling pathway in human B-lymphocyte precursors. Proc.

Natl. Acad. Sci. U. S. A. 90:252-6

192. Wilson RE, Taylor SL, Atherton GT, Johnston D, Waters CM, Norton JD

1993 Early response gene signalling cascades activated by ionising radiation in

primary human B cells. Oncogene 8:3229-37

157 193. Schreck R, Albermann K, Baeuerle PA 1992 Nuclear factor kappa B: an

oxidative stress-responsive transcription factor of eukaryotic cells (a review). Free

Radic. Res. Commun. 17:221-37

194. Mohan N, Meltz ML 1994 Induction of nuclear factor kB after low-dose ionizing

radiation involves a reactive oxygen intermediate signaling pathway. Radiat. Res.

140:97-104

195. Baeuml H, Behrends U, Peter RU, et al. 1997 Ionizing radiation induces, via

generation of reactive oxygen intermediates, intercellular adhesion molecule-1

(ICAM-1) gene transcription and NF-kB-like binding activity in the ICAM-1

transcriptional regulatory region. Free Radic. Res. 27:127-42

196. Iwanaga M, Mori K, Iida T, et al. 1998 Nuclear factor kappa B dependent

induction of gamma glutamylcysteine synthetase by ionizing radiation in T98G

human glioblastoma cells. Free Radic. Biol. Med. 24:1256-68

197. Li B, Tournier C, Davis RJ, Flavell RA 1999 Regulation of IL-4 expression by

the transcription factor JunB during T helper cell differentiation. Embo J 18:420-

32.

198. Schoonbroodt S, Piette J 2000 Oxidative stress interference with the nuclear

factor-kB activation pathways. Biochem. Pharmacol. 60:1075-83

199. Janssen YM, Van Houten B, Borm PJ, Mossman BT 1993 Cell and tissue

responses to oxidative damage. Lab Invest 69:261-74

200. Janssen-Heininger YM, Poynter ME, Baeuerle PA 2000 Recent advances

towards understanding redox mechanisms in the activation of nuclear factor kB.

Free. Radic. Biol. Med. 28:1317-27

158 201. Pajonk F, Riess K, Sommer A, McBride WH 2002 N-acetyl-L-cysteine inhibits

26S proteasome function: implications for effects on NF-kB activation. Free

Radic. Biol. Med. 32:536-43

202. Russell JS, Raju U, Gumin GJ, et al. 2002 Inhibition of radiation-induced

nuclear factor-kappaB activation by an anti-Ras single-chain antibody fragment:

lack of involvement in radiosensitization. Cancer Res. 62:2318-26

203. Finco TS, Westwick JK, Norris JL, Beg AA, Der CJ, Baldwin ASJ 1997

Oncogenic Ha-Ras-induced signaling activates NF-kB transcriptional activity,

which is required for cellular transformation. J. Biol. Chem. 272:24113-6

204. Abeyama K, Kawano K, Nakajima T, Takasaki I, Kitajima I, Maruyama I

1995 Interleukin 6 mediated differentiation and rescue of cell redox in PC12 cells

exposed to ionizing radiation. FEBS Lett. 364:298-300

205. Beetz A, Peter RU, Oppel T, et al. 2000 NF-kB and AP-1 are responsible for

inducibility of the IL-6 promoter by ionizing radiation in HeLa cells. Int. J.

Radiat. Biol. 76:1443-53

206. Zhou D, Yu T, Chen G, et al. 2001 Effects of NF-kappaB1 (p50) targeted gene

disruption on ionizing radiation-induced NF-kappaB activation and TNFalpha,

IL-1alpha, IL-1beta and IL-6 mRNA expression in vivo. Int J Radiat Biol 77:763-

72

207. Hallahan DE, Virudachalam S, Kuchibhotla J 1998 Nuclear factor kappaB

dominant negative genetic constructs inhibit X-ray induction of cell adhesion

molecules in the vascular endothelium. Cancer Res. 58:5484-8

159 208. Hallahan DE, Dunphy E, Virudachalam S, Sukhatme VP, Kufe DW,

Weichselbaum RR 1995 C-jun and Egr-1 participate in DNA synthesis and cell

survival in response to ionizing radiation exposure. J Biol Chem 270:30303-9

209. Jung M, Zhang Y, Dimtchev A, Dritschilo A 1998 Impaired regulation of

nuclear factor-kB results in apoptosis induced by gamma radiation. Radiat. Res.

149:596-601

210. Wang CY, Mayo MW, Baldwin ASJ 1996 TNF- and cancer therapy-induced

apoptosis: potentiation by inhibition of NF-kappaB. Science 274:784-7

211. Russo SM, Tepper JE, Baldwin ASJ, et al. 2001 Enhancement of

radiosensitivity by proteasome inhibition: implications for a role of NF-kappaB.

Int. J. Radiat. Oncol. Biol. Phys. 50:183-93

212. Yamagishi N, Miyakoshi J, Takebe H 1997 Enhanced radiosensitivity by

inhibition of nuclear factor kB activation in human malignant glioma cells. Int. J.

Radiat. Biol. 72:157-62

213. Honda N, Yagi K, Ding GR, Miyakoshi J 2002 Radiosensitization by

overexpression of the nonphosphorylation form of IkBa in human glioma cells. J.

Radiat. Res. 43:283-92

214. Kawai H, Yamada Y, Tatsuka M, Niwa O, Yamamoto K, Suzuki F 1999

Down-regulation of nuclear factor kB is required for p53-dependent apoptosis in

X-ray-irradiated mouse lymphoma cells and thymocytes. Cancer Res. 59:6038-41

215. Kato T, Duffey DC, Ondrey FG, et al. 2000 Cisplatin and radiation sensitivity

in human head and neck squamous carcinomas are independently modulated by

glutathione and transcription factor NF-kB. Head Neck 22:748-59

160 216. Didelot C, Barberi-Heyob M, Bianchi A, et al. 2001 Constitutive NF-kB

activity influences basal apoptosis and radiosensitivity of head-and-neck

carcinoma cell lines. Int. J. Radiat. Oncol. Biol. Phys. 51:1354-60

217. Eichholtz-Wirth H, Sagan D 2000 IkappaB/NF-kappaB mediated cisplatin

resistance in HeLa cells after low-dose gamma-irradiation is associated with

altered SODD expression. Apoptosis 5:255-63

218. Chen X, Shen B, Xia L, et al. 2002 Activation of nuclear factor kappaB in

radioresistance of TP53-inactive human keratinocytes. Cancer Res 62:1213-21

219. Pajonk F, Pajonk K, McBride WH 1999 Inhibition of NF-kB, clonogenicity,

and radiosensitivity of human cancer cells. J. Natl. Cancer. Inst. 91:1956-60

220. Pajonk F, Pajonk K, McBride WH 2000 Apoptosis and radiosensitization of

hodgkin cells by proteasome inhibition. Int. J. Radiat. Oncol. Biol. Phys. 47:1025-

32

221. Grana TM, Rusyn EV, Zhou H, Sartor CI, Cox AD 2002 Ras mediates

radioresistance through both phosphatidylinositol 3-kinase-dependent and Raf-

dependent but mitogen-activated protein kinase/extracellular signal-regulated

kinase kinase-independent signaling pathways. Cancer Res. 62:4142-50

222. Carlsen H, Moskaug JO, Fromm SH, Blomhoff R 2002 In vivo imaging of NF-

kB activity. J. Immunol. 168:1441-6

223. Durand RE, Olive PL 2001 Resistance of tumor cells to chemo- and

radiotherapy modulated by the three-dimensional architecture of solid tumors and

spheroids. Methods Cell Biol. 64:211-33

161 224. Weaver VM, Lelievre S, Lakins JN, et al. 2002 beta4 integrin-dependent

formation of polarized three-dimensional architecture confers resistance to

apoptosis in normal and malignant mammary epithelium. Cancer Cell 2:205-16

225. Baldwin ASJ 2001 Control of oncogenesis and cancer therapy resistance by the

transcription factor NF-kappaB. J. Clin. Invest. 107:241-6

226. Orlowski RZ, Baldwin ASJ 2002 NF-kB as a therapeutic target in cancer.

Trends Mol. Med. 8:385-9

227. Orlowski RZ, Stinchcombe TE, Mitchell BS, et al. 2002 Phase I Trial of the

Proteasome Inhibitor PS-341 in Patients With Refractory Hematologic

Malignancies. J. Clin. Oncol. 20:4420-7

228. Bohmann D, Admon A, Turner DR, Tjian R 1988 Transcriptional regulation

by the AP-1 family of enhancer-binding proteins: a nuclear target for signal

transduction. Cold Spring Harb Symp Quant Biol 53:695-700.

229. Lee W, Mitchell P, Tjian R 1987 Purified transcription factor AP-1 interacts

with TPA-inducible enhancer elements. Cell 49:741-52.

230. Cohen DR, Ferreira PC, Gentz R, Franza BR, Jr., Curran T 1989 The

product of a fos-related gene, fra-1, binds cooperatively to the AP- 1 site with Jun:

transcription factor AP-1 is comprised of multiple protein complexes. Genes Dev

3:173-84.

231. Hirai SI, Ryseck RP, Mechta F, Bravo R, Yaniv M 1989 Characterization of

junD: a new member of the jun proto-oncogene family. Embo J 8:1433-9.

162 232. Nishina H, Sato H, Suzuki T, Sato M, Iba H 1990 Isolation and characterization

of fra-2, an additional member of the fos gene family. Proc Natl Acad Sci U S A

87:3619-23.

233. Suzuki T, Okuno H, Yoshida T, Endo T, Nishina H, Iba H 1991 Difference in

transcriptional regulatory function between c-Fos and Fra- 2. Nucleic Acids Res

19:5537-42.

234. Wilkinson DG, Bhatt S, Ryseck RP, Bravo R 1989 Tissue-specific expression

of c-jun and junB during organogenesis in the mouse. Development 106:465-71.

235. Redemann-Fibi B, Schuermann M, Muller R 1991 Stage and tissue-specific

expression of fosB during mouse development. Differentiation 46:43-9.

236. De Cesare D, Vallone D, Caracciolo A, Sassone-Corsi P, Nerlov C, Verde P

1995 Heterodimerization of c-Jun with ATF-2 and c-Fos is required for positive

and negative regulation of the human urokinase enhancer. Oncogene 11:365-76.

237. Morooka H, Bonventre JV, Pombo CM, Kyriakis JM, Force T 1995 Ischemia

and reperfusion enhance ATF-2 and c-Jun binding to cAMP response elements

and to an AP-1 binding site from the c-jun promoter. J Biol Chem 270:30084-92.

238. Novotny V, Prieschl EE, Csonga R, Fabjani G, Baumruker T 1998 Nrf1 in a

complex with fosB, c-jun, junD and ATF2 forms the AP1 component at the TNF

alpha promoter in stimulated mast cells. Nucleic Acids Res 26:5480-5.

239. van Dam H, Duyndam M, Rottier R, et al. 1993 Heterodimer formation of cJun

and ATF-2 is responsible for induction of c-jun by the 243 amino acid adenovirus

E1A protein. Embo J 12:479-87.

163 240. van Dam H, Wilhelm D, Herr I, Steffen A, Herrlich P, Angel P 1995 ATF-2 is

preferentially activated by stress-activated protein kinases to mediate c-jun

induction in response to genotoxic agents. Embo J 14:1798-811.

241. Kataoka K, Noda M, Nishizawa M 1994 Maf nuclear oncoprotein recognizes

sequences related to an AP-1 site and forms heterodimers with both Fos and Jun.

Mol Cell Biol 14:700-12.

242. Kerppola TK, Curran T 1994 Maf and Nrl can bind to AP-1 sites and form

heterodimers with Fos and Jun. Oncogene 9:675-84.

243. Maki Y, Bos TJ, Davis C, Starbuck M, Vogt PK 1987 Avian sarcoma virus 17

carries the jun oncogene. Proc Natl Acad Sci U S A 84:2848-52.

244. Ryseck RP, Bravo R 1991 c-JUN, JUN B, and JUN D differ in their binding

affinities to AP-1 and CRE consensus sequences: effect of FOS proteins.

Oncogene 6:533-42.

245. Hai T, Curran T 1991 Cross-family dimerization of transcription factors Fos/Jun

and ATF/CREB alters DNA binding specificity. Proc Natl Acad Sci U S A

88:3720-4.

246. Kobierski LA, Chu HM, Tan Y, Comb MJ 1991 cAMP-dependent regulation

of proenkephalin by JunD and JunB: positive and negative effects of AP-1

proteins. Proc Natl Acad Sci U S A 88:10222-6.

247. Finch S, Joseloff E, Bowden T 2002 JunB negatively regulates AP-1 activity and

cell proliferation of malignant mouse keratinocytes. J Cancer Res Clin Oncol

128:3-10.

164 248. Adler V, Franklin CC, Kraft AS 1992 Phorbol esters stimulate the

phosphorylation of c-Jun but not v-Jun: regulation by the N-terminal delta

domain. Proc Natl Acad Sci U S A 89:5341-5.

249. Kyriakis JM, Banerjee P, Nikolakaki E, et al. 1994 The stress-activated protein

kinase subfamily of c-Jun kinases. Nature 369:156-60.

250. Giardina SL, Evans SW, Gandino L, et al. 1987 Generation of a murine

monoclonal antibody that detects the fos oncogene product. Anal Biochem

161:109-16.

251. Ransone LJ, Wamsley P, Morley KL, Verma IM 1990 Domain swapping

reveals the modular nature of Fos, Jun, and CREB proteins. Mol Cell Biol

10:4565-73.

252. Nel AE, Taylor LK, Kumar GP, et al. 1994 Activation of a novel

serine/threonine kinase that phosphorylates c-Fos upon stimulation of T and B

lymphocytes via antigen and cytokine receptors. J Immunol 152:4347-57.

253. Taylor LK, Swanson KD, Kerigan J, Mobley W, Landreth GE 1994 Isolation

and characterization of a nerve growth factor-regulated Fos kinase from PC12

cells. J Biol Chem 269:308-18.

254. Masquilier D, Sassone-Corsi P 1992 Transcriptional cross-talk: nuclear factors

CREM and CREB bind to AP-1 sites and inhibit activation by Jun. J Biol Chem

267:22460-6.

255. Auble DT, Brinckerhoff CE 1991 The AP-1 sequence is necessary but not

sufficient for phorbol induction of collagenase in fibroblasts. Biochemistry

30:4629-35.

165 256. Kick G, Messer G, Goetz A, Plewig G, Kind P 1995 Photodynamic therapy

induces expression of interleukin 6 by activation of AP-1 but not NF-kappa B

DNA binding. Cancer Res 55:2373-9.

257. Higo H, Lee JY, Satow Y, Higo K 1989 Elevated expression of proto-oncogenes

accompany enhanced induction of heat-shock genes after exposure of rat embryos

in utero to ionizing irradiation. Teratog Carcinog Mutagen 9:191-8

258. Woloschak GE, Chang-Liu CM 1990 Differential modulation of specific gene

expression following high- and low-LET radiations. Radiat Res 124:183-7.

259. Datta R, Hallahan DE, Kharbanda SM, et al. 1992 Involvement of reactive

oxygen intermediates in the induction of c-jun gene transcription by ionizing

radiation. Biochemistry 31:8300-6.

260. Datta R, Hass R, Gunji H, Weichselbaum R, Kufe D 1992 Down-regulation of

cell cycle control genes by ionizing radiation. Cell Growth Differ 3:637-44.

261. Ishihara H, Tsuneoka K, Dimchev AB, Shikita M 1993 Induction of the

expression of the interleukin-1 beta gene in mouse spleen by ionizing radiation.

Radiat Res 133:321-6.

262. Manome Y, Datta R, Fine HA 1993 Early response gene induction following

DNA damage in astrocytoma cell lines. Biochem Pharmacol 45:1677-84

263. Prasad AV, Mohan N, Chandrasekar B, Meltz ML 1995 Induction of

transcription of "immediate early genes" by low-dose ionizing radiation. Radiat

Res 143:263-72.

264. Hallahan DE 1996 Radiation-Mediated Gene Expression in the Pathogenesis of

the Clinical Radiation Response. Semin Radiat Oncol 6:250-267

166 265. Minet E, Michel G, Mottet D, et al. 2001 c-JUN gene induction and AP-1

activity is regulated by a JNK-dependent pathway in hypoxic HepG2 cells. Exp

Cell Res 265:114-24.

266. Hallahan DE, Virudachalam S, Beckett M, Sherman ML, Kufe D,

Weichselbaum RR 1991 Mechanisms of X-ray-mediated protooncogene c-jun

expression in radiation-induced human sarcoma cell lines. Int J Radiat Oncol Biol

Phys 21:1677-81.

267. Zhou H, Lin A, Gu Z, Chen S, Park NH, Chiu R 2000 12-O-

tetradecanoylphorbol-13-acetate (TPA)-induced c-Jun N-terminal kinase (JNK)

phosphatase renders immortalized or transformed epithelial cells refractory to

TPA-inducible JNK activity. J Biol Chem 275:22868-75.

268. Choi EK, Rhee YH, Park HJ, Ahn SD, Shin KH, Park KK 2001 Effect of

protein kinase C inhibitor (PKCI) on radiation sensitivity and c-fos transcription.

Int J Radiat Oncol Biol Phys 49:397-405.

269. Lee SA, Dritschilo A, Jung M 2001 Role of ATM in oxidative stress-mediated

c-Jun phosphorylation in response to ionizing radiation and CdCl2. J Biol Chem

276:11783-90.

270. Verheij M, Ruiter GA, Zerp SF, et al. 1998 The role of the stress-activated

protein kinase (SAPK/JNK) signaling pathway in radiation-induced apoptosis.

Radiother Oncol 47:225-32.

271. Martin M, Vozenin MC, Gault N, Crechet F, Pfarr CM, Lefaix JL 1997

Coactivation of AP-1 activity and TGF-beta1 gene expression in the stress

response of normal skin cells to ionizing radiation. Oncogene 15:981-9.

167 272. Park JS, Qiao L, Su ZZ, et al. 2001 Ionizing radiation modulates vascular

endothelial growth factor (VEGF) expression through multiple mitogen activated

protein kinase dependent pathways. Oncogene 20:3266-80

273. Gault N, Vozenin-Brotons MC, Calenda A, Lefaix JL, Martin MT 2002

Promoter sequences involved in transforming growth factor beta1 gene induction

in HaCaT keratinocytes after gamma irradiation. Radiat Res 157:249-55.

274. Black AR, Black JD, Azizkhan-Clifford J 2001 Sp1 and kruppel-like factor

family of transcription factors in cell growth regulation and cancer. J Cell Physiol

188:143-60

275. Nenoi M, Ichimura S, Mita K, Yukawa O, Cartwright IL 2001 Regulation of

the catalase gene promoter by Sp1, CCAAT-recognizing factors, and a WT1/Egr-

related factor in hydrogen peroxide-resistant HP100 cells. Cancer Res 61:5885-94

276. Datta R, Rubin E, Sukhatme V, et al. 1992 Ionizing radiation activates

transcription of the EGR1 gene via CArG elements. Proc Natl Acad Sci U S A

89:10149-53

277. Cao XM, Koski RA, Gashler A, et al. 1990 Identification and characterization

of the Egr-1 gene product, a DNA-binding zinc finger protein induced by

differentiation and growth signals. Mol Cell Biol 10:1931-9

278. Cao XM, Guy GR, Sukhatme VP, Tan YH 1992 Regulation of the Egr-1 gene

by tumor necrosis factor and interferons in primary human fibroblasts. J Biol

Chem 267:1345-9

279. Vrana KE, Churchill ME, Tullius TD, Brown DD 1988 Mapping functional

regions of transcription factor TFIIIA. Mol Cell Biol 8:1684-96

168 280. Stillman DJ, Bankier AT, Seddon A, Groenhout EG, Nasmyth KA 1988

Characterization of a transcription factor involved in mother cell specific

transcription of the yeast HO gene. Embo J 7:485-94

281. Qureshi SA, Cao XM, Sukhatme VP, Foster DA 1991 v-Src activates mitogen-

responsive transcription factor Egr-1 via serum response elements. J Biol Chem

266:10802-6

282. Alexandropoulos K, Qureshi SA, Rim M, Sukhatme VP, Foster DA 1992 v-

Fps-responsiveness in the Egr-1 promoter is mediated by serum response

elements. Nucleic Acids Res 20:2355-9

283. Ahmed MM, Sells SF, Venkatasubbarao K, et al. 1997 Ionizing radiation-

inducible apoptosis in the absence of p53 linked to transcription factor EGR-1. J

Biol Chem 272:33056-61

284. Ahmed MM, Venkatasubbarao K, Fruitwala SM, et al. 1996 EGR-1 induction

is required for maximal radiosensitivity in A375-C6 melanoma cells. J Biol Chem

271:29231-7

285. Takahashi T, Namiki Y, Ohno T 1997 Induction of the suicide HSV-TK gene

by activation of the Egr-1 promoter with radioisotopes. Hum Gene Ther 8:827-33

286. Meyer RG, Kupper JH, Kandolf R, Rodemann HP 2002 Early growth

response-1 gene (Egr-1) promoter induction by ionizing radiation in U87

malignant glioma cells in vitro. Eur J Biochem 269:337-46

287. Weichselbaum RR, Kufe DW, Hellman S, et al. 2002 Radiation-induced

tumour necrosis factor-alpha expression: clinical application of transcriptional and

physical targeting of gene therapy. Lancet Oncol 3:665-71

169 288. Weichselbaum RR, Hallahan D, Fuks Z, Kufe D 1994 Radiation induction of

immediate early genes: effectors of the radiation-stress response. Int J Radiat

Oncol Biol Phys 30:229-34.

289. Hallahan DE, Sukhatme VP, Sherman ML, Virudachalam S, Kufe D,

Weichselbaum RR 1991 Protein kinase C mediates x-ray inducibility of nuclear

signal transducers EGR1 and JUN. Proc Natl Acad Sci U S A 88:2156-60

290. Hallahan DE, Mauceri HJ, Seung LP, et al. 1995 Spatial and temporal control

of gene therapy using ionizing radiation. Nat Med 1:786-91

291. Hallahan D, Clark ET, Kuchibhotla J, Gewertz BL, Collins T 1995 E-selectin

gene induction by ionizing radiation is independent of cytokine induction.

Biochem. Biophys. Res. Commun. 217:784-95

292. Zhao H, Jin S, Fan F, Fan W, Tong T, Zhan Q 2000 Activation of the

transcription factor Oct-1 in response to DNA damage. Cancer Res 60:6276-80

293. Jin S, Fan F, Fan W, et al. 2001 Transcription factors Oct-1 and NF-YA regulate

the p53-independent induction of the GADD45 following DNA damage.

Oncogene 20:2683-90

294. Bertagna A, Jahroudi N 2001 The NFY transcription factor mediates induction

of the von Willebrand factor promoter by irradiation. Thromb Haemost 85:837-44

295. Reddy KB, Jin G, Karode MC, Harmony JA, Howe PH 1996 Transforming

growth factor beta (TGF beta)-induced nuclear localization of apolipoprotein

J/clusterin in epithelial cells. Biochemistry 35:6157-63

296. Smith ML, Zhan Q, Bae I, Fornace AJ, Jr. 1994 Role of retinoblastoma gene

product in p53-mediated DNA damage response. Exp Cell Res 215:386-9

170 297. Sheikh MS, Fornace AJ, Jr. 2000 Role of p53 family members in apoptosis. J

Cell Physiol 182:171-81

298. Sheikh MS, Hollander MC, Fornance AJ, Jr. 2000 Role of Gadd45 in

apoptosis. Biochem Pharmacol 59:43-5

299. Goldwasser F, Bae I, Pommier Y, Fornace AJ, Jr. 1999 Evidence of a reduced

DNA topoisomerase II mRNA expression after ionizing radiation. Anticancer Res

19:3167-71

300. Fornace AJ, Jr., Amundson SA, Bittner M, et al. 1999 The complexity of

radiation stress responses: analysis by informatics and functional genomics

approaches. Gene Expr 7:387-400

301. Fornace AJ, Jr., Amundson SA, Do KT, Meltzer P, Trent J, Bittner M 2002

Stress-gene induction by low-dose gamma irradiation. Military Medicine 167:13-

5

302. Bae I, Smith ML, Sheikh MS, et al. 1996 An abnormality in the p53 pathway

following gamma-irradiation in many wild-type p53 human melanoma lines.

Cancer Res 56:840-7

303. Bae I, Fan S, Bhatia K, Kohn KW, Fornace AJ, Jr., O'Connor PM 1995

Relationships between G1 arrest and stability of the p53 and p21Cip1/Waf1

proteins following gamma-irradiation of human lymphoma cells. Cancer Res

55:2387-93

304. Ueda T, Akiyama N, Sai H, et al. 2001 c-IAP2 is induced by ionizing radiation

through NF-kB binding sites. FEBS Lett. 491:40-4

171 305. McDonald JF, Nelsestuen GL 1997 Potent inhibition of terminal complement

assembly by clusterin: characterization of its impact on C9 polymerization.

Biochemistry 36:7464-73

306. Polihronis M, Paizis K, Carter G, Sedal L, Murphy B 1993 Elevation of

human cerebrospinal fluid clusterin concentration is associated with acute

neuropathology. J Neurol Sci 115:230-3.

307. Urbich C, Fritzenwanger M, Zeiher AM, Dimmeler S 2000 Laminar shear

stress upregulates the complement-inhibitory protein clusterin : a novel potent

defense mechanism against complement-induced endothelial cell activation.

Circulation 101:352-5

308. Ishikawa Y, Akasaka Y, Ishii T, et al. 1998 Distribution and synthesis of

apolipoprotein J in the atherosclerotic aorta. Arterioscler Thromb Vasc Biol

18:665-72.

309. DeMattos RB, Brendza RP, Heuser JE, et al. 2001 Purification and

characterization of astrocyte-secreted apolipoprotein E and J-containing

lipoproteins from wild-type and human apoE transgenic mice. Neurochem Int

39:415-25.

310. Miyake H, Nelson C, Rennie PS, Gleave ME 2000 Acquisition of

chemoresistant phenotype by overexpression of the antiapoptotic gene

testosterone-repressed prostate message-2 in prostate cancer xenograft models.

Cancer Res 60:2547-54

311. Levine AJ, Momand J, Finlay CA 1991 The p53 tumour suppressor gene.

Nature 351:453-6

172 312. Canman CE, Chen CY, Lee MH, Kastan MB 1994 DNA damage responses:

p53 induction, cell cycle perturbations, and apoptosis. Cold Spring Harb Symp

Quant Biol 59:277-86

313. Yonish-Rouach E, Grunwald D, Wilder S, et al. 1993 p53-mediated cell death:

relationship to cell cycle control. Mol Cell Biol 13:1415-23

314. Kastan MB, Canman CE, Leonard CJ 1995 P53, cell cycle control and

apoptosis: implications for cancer. Cancer Metastasis Rev 14:3-15

315. Roperch JP, Alvaro V, Prieur S, et al. 1998 Inhibition of presenilin 1

expression is promoted by p53 and p21WAF-1 and results in apoptosis and tumor

suppression. Nat Med 4:835-8

316. Agoff SN, Hou J, Linzer DI, Wu B 1993 Regulation of the human hsp70

promoter by p53. Science 259:84-7

317. Yamamoto M, Yoshida M, Ono K, et al. 1994 Effect of tumor suppressors on

cell cycle-regulatory genes: RB suppresses p34cdc2 expression and normal p53

suppresses cyclin A expression. Exp Cell Res 210:94-101.

318. Krause K, Wasner M, Reinhard W, et al. 2000 The tumour suppressor protein

p53 can repress transcription of cyclin B. Nucleic Acids Res 28:4410-8.

319. Taylor WR, Schonthal AH, Galante J, Stark GR 2001 p130/E2F4 binds to and

represses the cdc2 promoter in response to p53. J Biol Chem 276:1998-2006.

320. Akakura K, Bruchovsky N, Rennie PS, et al. 1996 Effects of intermittent

androgen suppression on the stem cell composition and the expression of the

TRPM-2 (clusterin) gene in the Shionogi carcinoma. J Steroid Biochem Mol Biol

59:501-11

173 321. Bunz F, Dutriaux A, Lengauer C, et al. 1998 Requirement for p53 and p21 to

sustain G2 arrest after DNA damage. Science 282:1497-501

322. Pink JJ, Wuerzberger-Davis S, Tagliarino C, et al. 2000 Activation of a

cysteine protease in MCF-7 and T47D breast cancer cells during beta-lapachone-

mediated apoptosis. Exp Cell Res 255:144-55.

323. Agani FH, Puchowicz M, Chavez JC, Pichiule P, LaManna J 2002 Role of

nitric oxide in the regulation of HIF-1alpha expression during hypoxia. Am J

Physiol Cell Physiol 283:C178-86

324. Wuerzberger SM, Pink JJ, Planchon SM, Byers KL, Bornmann WG,

Boothman DA 1998 Induction of apoptosis in MCF-7:WS8 breast cancer cells by

beta-lapachone. Cancer Res 58:1876-85

325. Bach UC, Baiersdorfer M, Klock G, Cattaruzza M, Post A, Koch-Brandt C

2001 Apoptotic cell debris and phosphatidylserine-containing lipid vesicles

induce apolipoprotein J (clusterin) gene expression in vital fibroblasts. Exp Cell

Res 265:11-20.

326. Pink JJ, Planchon SM, Tagliarino C, Varnes ME, Siegel D, Boothman DA

2000 NAD(P)H:Quinone oxidoreductase activity is the principal determinant of

beta-lapachone cytotoxicity. J Biol Chem 275:5416-24

327. Scheffner M, Werness BA, Huibregtse JM, Levine AJ, Howley PM 1990 The

E6 oncoprotein encoded by human papillomavirus types 16 and 18 promotes the

degradation of p53. Cell 63:1129-36.

174 328. Bettuzzi S, Astancolle S, Guidetti G, Moretti M, Tiozzo R, Corti A 1999

Clusterin (SGP-2) gene expression is cell cycle dependent in normal human

dermal fibroblasts. FEBS Lett 448:297-300

329. Meyers M, Theodosiou M, Acharya S, et al. 1997 Cell cycle regulation of the

human DNA mismatch repair genes hMSH2, hMLH1, and hPMS2. Cancer Res

57:206-8

330. Meyers M, Wagner MW, Hwang HS, Kinsella TJ, Boothman DA 2001 Role

of the hMLH1 DNA mismatch repair protein in fluoropyrimidine-mediated cell

death and cell cycle responses. Cancer Res 61:5193-201.

331. Tagliarino C, Pink JJ, Dubyak GR, Nieminen AL, Boothman DA 2001

Calcium is a key signaling molecule in beta-lapachone-mediated cell death. J Biol

Chem 276:19150-9.

332. Budhram-Mahadeo V, Morris PJ, Smith MD, Midgley CA, Boxer LM,

Latchman DS 1999 p53 suppresses the activation of the Bcl-2 promoter by the

Brn-3a POU family transcription factor. J Biol Chem 274:15237-44.

333. Lee KC, Crowe AJ, Barton MC 1999 p53-mediated repression of alpha-

fetoprotein gene expression by specific DNA binding. Mol Cell Biol 19:1279-88.

334. Ori A, Zauberman A, Doitsh G, Paran N, Oren M, Shaul Y 1998 p53 binds

and represses the HBV enhancer: an adjacent enhancer element can reverse the

transcription effect of p53. Embo J 17:544-53.

335. Kanaya T, Kyo S, Hamada K, et al. 2000 Adenoviral expression of p53

represses telomerase activity through down-regulation of human telomerase

reverse transcriptase transcription. Clin Cancer Res 6:1239-47.

175 336. Ohlsson C, Kley N, Werner H, LeRoith D 1998 p53 regulates insulin-like

growth factor-I (IGF-I) receptor expression and IGF-I-induced tyrosine

phosphorylation in an osteosarcoma cell line: interaction between p53 and Sp1.

Endocrinology 139:1101-7.

337. Sun Y, Wenger L, Rutter JL, Brinckerhoff CE, Cheung HS 1999 p53 down-

regulates human matrix metalloproteinase-1 (Collagenase-1) gene expression. J

Biol Chem 274:11535-40.

338. Subbaramaiah K, Altorki N, Chung WJ, Mestre JR, Sampat A, Dannenberg

AJ 1999 Inhibition of cyclooxygenase-2 gene expression by p53. J Biol Chem

274:10911-5.

339. Yun J, Chae HD, Choy HE, et al. 1999 p53 negatively regulates cdc2

transcription via the CCAAT-binding NF-Y transcription factor. J Biol Chem

274:29677-82.

340. Kubicka S, Kuhnel F, Zender L, et al. 1999 p53 represses CAAT enhancer-

binding protein (C/EBP)-dependent transcription of the albumin gene. A

molecular mechanism involved in viral liver infection with implications for

hepatocarcinogenesis. J Biol Chem 274:32137-44.

341. Martin DW, Munoz RM, Subler MA, Deb S 1993 p53 binds to the TATA-

binding protein-TATA complex. J Biol Chem 268:13062-7

342. Johnson RA, Ince TA, Scotto KW 2001 Transcriptional repression by p53

through direct binding to a novel DNA element. J Biol Chem 276:27716-20.

343. Schmidt-Ullrich RK, Valerie KC, Chan W, McWilliams D 1994 Altered

expression of epidermal growth factor receptor and estrogen receptor in MCF-7

176 cells after single and repeated radiation exposures. Int J Radiat Oncol Biol Phys

29:813-9

344. Schmidt-Ullrich RK, Mikkelsen RB, Dent P, et al. 1997 Radiation-induced

proliferation of the human A431 squamous carcinoma cells is dependent on

EGFR tyrosine phosphorylation. Oncogene 15:1191-7

345. Kim KU, Xiao J, Ni HT, et al. 2003 Changes in expression of transferrin,

insulin-like growth factor 1, and interleukin 4 receptors after irradiation of cells of

primary malignant brain tumor cell lines. Radiat Res 160:224-31

346. Leverrier Y, Thomas J, Mathieu AL, Low W, Blanquier B, Marvel J 1999

Role of PI3-kinase in Bcl-X induction and apoptosis inhibition mediated by IL-3

or IGF-1 in Baf-3 cells. Cell Death Differ 6:290-6

347. Sherman ML, Datta R, Hallahan DE, Weichselbaum RR, Kufe DW 1991

Regulation of tumor necrosis factor gene expression by ionizing radiation in

human myeloid leukemia cells and peripheral blood monocytes. J Clin Invest

87:1794-7

348. Dent P, Yacoub A, Contessa J, et al. 2003 Stress and radiation-induced

activation of multiple intracellular signaling pathways. Radiat Res 159:283-300

349. Dent P, Reardon DB, Park JS, et al. 1999 Radiation-induced release of

transforming growth factor alpha activates the epidermal growth factor receptor

and mitogen-activated protein kinase pathway in carcinoma cells, leading to

increased proliferation and protection from radiation-induced cell death. Mol Biol

Cell 10:2493-506

177 350. Eichholtz-Wirth H, Sagan D 2002 Altered signaling of TNFalpha-TNFR1 and

SODD/BAG4 is responsible for radioresistance in human HT-R15 cells.

Anticancer Res 22:235-40

351. Grana TM, Sartor CI, Cox AD 2003 Epidermal growth factor receptor

autocrine signaling in RIE-1 cells transformed by the Ras oncogene enhances

radiation resistance. Cancer Res 63:7807-14

352. Yu D, Watanabe H, Shibuya H, Miura M 2003 Redundancy of radioresistant

signaling pathways originating from insulin-like growth factor I receptor. J Biol

Chem 278:6702-9

353. Vogel CL, Franco SX 2003 Clinical experience with trastuzumab (herceptin).

Breast J 9:452-62

354. Tan AR, Swain SM 2003 Ongoing adjuvant trials with trastuzumab in breast

cancer. Semin Oncol 30:54-64

355. Arteaga CL 2003 Inhibiting tyrosine kinases: successes and limitations. Cancer

Biol Ther 2:S79-83

356. De Marinis F, Nelli F, D'Auria G 2002 EGFR inhibitors: clinical results. Suppl

Tumori 1:S5-6

357. Grunwald V, Hidalgo M 2003 Developing inhibitors of the epidermal growth

factor receptor for cancer treatment. J Natl Cancer Inst 95:851-67

358. Bartl MM, Luckenbach T, Bergner O, Ullrich O, Koch-Brandt C 2001

Multiple receptors mediate apoJ-dependent clearance of cellular debris into

nonprofessional phagocytes. Exp Cell Res 271:130-41.

178 359. Jenne DE, Tschopp J 1992 Clusterin: the intriguing guises of a widely expressed

glycoprotein. Trends Biochem Sci 17:154-9

360. Brown TL, Moulton BC, Baker VV, Mira J, Harmony JA 1995 Expression of

apolipoprotein J in the uterus is associated with tissue remodeling. Biol Reprod

52:1038-49

361. Atlas-White M, Murphy BF, Baker HW 2000 Localisation of clusterin in

normal human sperm by immunogold electron microscopy. Pathology 32:258-61.

362. Wong P, Taillefer D, Lakins J, Pineault J, Chader G, Tenniswood M 1994

Molecular characterization of human TRPM-2/clusterin, a gene associated with

sperm maturation, apoptosis and . Eur J Biochem 221:917-25

363. Trougakos IP, Gonos ES 2002 Clusterin/apolipoprotein J in human aging and

cancer. Int J Biochem Cell Biol 34:1430-48

364. Yamada K, Hori Y, Hanafusa N, et al. 2001 Clusterin is up-regulated in

glomerular mesangial cells in complement-mediated injury. Kidney Int 59:137-

46.

365. Burkey BF, Stuart WD, Harmony JA 1992 Hepatic apolipoprotein J is secreted

as a lipoprotein. J Lipid Res 33:1517-26.

366. DeMattos RB, O'Dell M A, Parsadanian M, et al. 2002 Clusterin promotes

amyloid plaque formation and is critical for neuritic toxicity in a mouse model of

Alzheimer's disease. Proc Natl Acad Sci U S A 99:10843-8

367. Harr SD, Uint L, Hollister R, Hyman BT, Mendez AJ 1996 Brain expression

of apolipoproteins E, J, and A-I in Alzheimer's disease. J Neurochem 66:2429-35.

179 368. Trougakos IP, Poulakou M, Stathatos M, Chalikia A, Melidonis A, Gonos ES

2002 Serum levels of the senescence biomarker clusterin/apolipoprotein J

increase significantly in diabetes type II and during development of coronary

heart disease or at . Exp Gerontol 37:1175-87

369. Jordan-Starck TC, Lund SD, Witte DP, et al. 1994 Mouse apolipoprotein J:

characterization of a gene implicated in atherosclerosis. J Lipid Res 35:194-210.

370. Chiesa R, Angeretti N, Lucca E, et al. 1996 Clusterin (SGP-2) induction in rat

astroglial cells exposed to prion protein fragment 106-126. Eur J Neurosci 8:589-

97

371. McHattie S, Wells GA, Bee J, Edington N 1999 Clusterin in bovine spongiform

encephalopathy (BSE). J Comp Pathol 121:159-71

372. Sasaki K, Doh-ura K, Ironside JW, Iwaki T 2002 Increased clusterin

(apolipoprotein J) expression in human and mouse brains infected with

transmissible spongiform encephalopathies. Acta Neuropathol (Berl) 103:199-208

373. Aigner T, Dudhia J 2003 Genomics of osteoarthritis. Curr Opin Rheumatol

15:634-40

374. Hidaka S, Kranzlin B, Gretz N, Witzgall R 2002 Urinary clusterin levels in the

rat correlate with the severity of tubular damage and may help to differentiate

between glomerular and tubular injuries. Cell Tissue Res 310:289-96

375. Ghiggeri GM, Bruschi M, Candiano G, et al. 2002 Depletion of clusterin in

renal diseases causing nephrotic syndrome. Kidney Int 62:2184-94

180 376. Zellweger T, Kiyama S, Chi K, et al. 2003 Overexpression of the cytoprotective

protein clusterin decreases radiosensitivity in the human LNCaP prostate tumour

model. BJU Int 92:463-9

377. Bettuzzi S 2003 The new anti-oncogene clusterin and the molecular profiling of

prostate cancer progression and prognosis. Acta Biomed Ateneo Parmense

74:101-4

378. Miyake H, Gleave ME, Arakawa S, Kamidono S, Hara I 2002 Introducing the

clusterin gene into human renal cell carcinoma cells enhances their metastatic

potential. J Urol 167:2203-8

379. July LV, Akbari M, Zellweger T, Jones EC, Goldenberg SL, Gleave ME 2002

Clusterin expression is significantly enhanced in prostate cancer cells following

androgen withdrawal therapy. Prostate 50:179-88

380. Zellweger T, Chi K, Miyake H, et al. 2002 Enhanced radiation sensitivity in

prostate cancer by inhibition of the cell survival protein clusterin. Clin Cancer Res

8:3276-84

381. Abbott DW, Holt JT 1999 Mitogen-activated protein kinase kinase 2 activation

is essential for progression through the G2/M checkpoint arrest in cells exposed to

ionizing radiation. J Biol Chem 274:2732-42

382. Schreiber E, Matthias P, Muller MM, Schaffner W 1989 Rapid detection of

octamer binding proteins with 'mini-extracts', prepared from a small number of

cells. Nucleic Acids Res 17:6419

383. Amorino GP, Hamilton VM, Valerie K, Dent P, Lammering G, Schmidt-

Ullrich RK 2002 Epidermal growth factor receptor dependence of radiation-

181 induced transcription factor activation in human breast carcinoma cells. Mol Biol

Cell 13:2233-44

384. Ahmad T, Farnie G, Bundred NJ, Anderson NG 2004 The Mitogenic Action

of Insulin-like Growth Factor I in Normal Human Mammary Epithelial Cells

Requires the Epidermal Growth Factor Receptor Tyrosine Kinase. J Biol Chem

279:1713-9

385. Wang D, Patil S, Li W, Humphrey LE, Brattain MG, Howell GM 2002

Activation of the TGFalpha autocrine loop is downstream of IGF-I receptor

activation during mitogenesis in growth factor dependent human colon carcinoma

cells. Oncogene 21:2785-96

386. Burgaud JL, Baserga R 1996 Intracellular transactivation of the insulin-like

growth factor I receptor by an epidermal growth factor receptor. Exp Cell Res

223:412-9

387. Freed KA, Herington AC 1989 Insulin-like growth factor-I and its autocrine role

in growth of MCF-7 human breast cancer cells in culture. J Mol Endocrinol

3:183-90

388. Gruden G, Araf S, Zonca S, et al. 2003 IGF-I induces vascular endothelial

growth factor in human mesangial cells via a Src-dependent mechanism. Kidney

Int 63:1249-55

389. Sekharam M, Nasir A, Kaiser HE, Coppola D 2003 Insulin-like growth factor

1 receptor activates c-SRC and modifies transformation and motility of colon

cancer in vitro. Anticancer Res 23:1517-24

182 390. Boney CM, Sekimoto H, Gruppuso PA, Frackelton AR, Jr. 2001 Src family

tyrosine kinases participate in insulin-like growth factor I mitogenic signaling in

3T3-L1 cells. Cell Growth Differ 12:379-86

391. Lebrun P, Mothe-Satney I, Delahaye L, Van Obberghen E, Baron V 1998

Insulin receptor substrate-1 as a signaling molecule for focal adhesion kinase

pp125(FAK) and pp60(src). J Biol Chem 273:32244-53

392. Kyriakis JM, App H, Zhang XF, et al. 1992 Raf-1 activates MAP kinase-

kinase. Nature 358:417-21

393. Jones N, Agani FH 2003 Hyperoxia induces Egr-1 expression through activation

of extracellular signal-regulated kinase 1/2 pathway. J Cell Physiol 196:326-33

394. Guha M, O'Connell MA, Pawlinski R, et al. 2001 Lipopolysaccharide

activation of the MEK-ERK1/2 pathway in human monocytic cells mediates

tissue factor and tumor necrosis factor alpha expression by inducing Elk-1

phosphorylation and Egr-1 expression. Blood 98:1429-39

395. Lo LW, Cheng JJ, Chiu JJ, Wung BS, Liu YC, Wang DL 2001 Endothelial

exposure to hypoxia induces Egr-1 expression involving PKCalpha-mediated

Ras/Raf-1/ERK1/2 pathway. J Cell Physiol 188:304-12

396. Hodge C, Liao J, Stofega M, Guan K, Carter-Su C, Schwartz J 1998 Growth

hormone stimulates phosphorylation and activation of elk-1 and expression of c-

fos, egr-1, and junB through activation of extracellular signal-regulated kinases 1

and 2. J Biol Chem 273:31327-36

183 397. Milas L, Fan Z, Andratschke NH, Ang KK 2004 Epidermal growth factor

receptor and tumor response to radiation: in vivo preclinical studies. Int J Radiat

Oncol Biol Phys 58:966-71

398. Krishnan S, Rao RD, James CD, Sarkaria JN 2003 Combination of epidermal

growth factor receptor targeted therapy with radiation therapy for malignant

gliomas. Front Biosci 8:e1-13

399. Zhou H, Kim YS, Peletier A, McCall W, Earp HS, Sartor CI 2004 Effects of

the EGFR/HER2 kinase inhibitor GW572016 on EGFR- and HER2-

overexpressing breast cancer cell line proliferation, radiosensitization, and

resistance. Int J Radiat Oncol Biol Phys 58:344-52

400. Gooch JL, Van Den Berg CL, Yee D 1999 Insulin-like growth factor (IGF)-I

rescues breast cancer cells from chemotherapy-induced cell death--proliferative

and anti-apoptotic effects. Breast Cancer Res Treat 56:1-10

401. Wen B, Deutsch E, Marangoni E, et al. 2001 Tyrphostin AG 1024 modulates

radiosensitivity in human breast cancer cells. Br J Cancer 85:2017-21

402. Dricu A, Kanter L, Wang M, et al. 1999 Expression of the insulin-like growth

factor 1 receptor (IGF-1R) in breast cancer cells: evidence for a regulatory role of

dolichyl phosphate in the transition from an intracellular to an extracellular IGF-1

pathway. Glycobiology 9:571-9

403. LeBedis C, Chen K, Fallavollita L, Boutros T, Brodt P 2002 Peripheral lymph

node stromal cells can promote growth and tumorigenicity of breast carcinoma

cells through the release of IGF-I and EGF. Int J Cancer 100:2-8

184 404. Bianco C, Tortora G, Bianco R, et al. 2002 Enhancement of antitumor activity

of ionizing radiation by combined treatment with the selective epidermal growth

factor receptor-tyrosine kinase inhibitor ZD1839 (Iressa). Clin Cancer Res

8:3250-8

405. Lu Y, Zi X, Zhao Y, Mascarenhas D, Pollak M 2001 Insulin-like growth factor-

I receptor signaling and resistance to trastuzumab (Herceptin). J Natl Cancer Inst

93:1852-7

406. Hatters DM, Wilson MR, Easterbrook-Smith SB, Howlett GJ 2002

Suppression of apolipoprotein C-II amyloid formation by the extracellular

chaperone, clusterin. Eur J Biochem 269:2789-94

407. Ho J, Benchimol S 2003 Transcriptional repression mediated by the p53 tumour

suppressor. Cell Death Differ 10:404-8

408. Murphy M, Ahn J, Walker KK, et al. 1999 Transcriptional repression by wild-

type p53 utilizes histone deacetylases, mediated by interaction with mSin3a.

Genes Dev 13:2490-501

409. Koumenis C, Alarcon R, Hammond E, et al. 2001 Regulation of p53 by

hypoxia: dissociation of transcriptional repression and apoptosis from p53-

dependent transactivation. Mol Cell Biol 21:1297-310

410. Mirza A, McGuirk M, Hockenberry TN, et al. 2002 Human survivin is

negatively regulated by wild-type p53 and participates in p53-dependent apoptotic

pathway. Oncogene 21:2613-22

411. Grossman SR, Deato ME, Brignone C, et al. 2003 Polyubiquitination of p53 by

a ubiquitin ligase activity of p300. Science 300:342-4

185 412. Lu Y, Yu Q, Liu JH, et al. 2003 Src family protein-tyrosine kinases alter the

function of PTEN to regulate phosphatidylinositol 3-kinase/AKT cascades. J Biol

Chem 278:40057-66

413. Weng LP, Smith WM, Brown JL, Eng C 2001 PTEN inhibits insulin-

stimulated MEK/MAPK activation and cell growth by blocking IRS-1

phosphorylation and IRS-1/Grb-2/Sos complex formation in a breast cancer

model. Hum Mol Genet 10:605-16

414. Tanno S, Mitsuuchi Y, Altomare DA, Xiao GH, Testa JR 2001 AKT activation

up-regulates insulin-like growth factor I receptor expression and promotes

invasiveness of human pancreatic cancer cells. Cancer Res 61:589-93

415. Manome Y, Kunieda T, Wen PY, Koga T, Kufe DW, Ohno T 1998 Transgene

expression in malignant glioma using a replication-defective adenoviral vector

containing the Egr-1 promoter: activation by ionizing radiation or uptake of

radioactive iododeoxyuridine. Hum Gene Ther 9:1409-17

416. Joki T, Nakamura M, Ohno T 1995 Activation of the radiosensitive EGR-1

promoter induces expression of the herpes simplex virus thymidine kinase gene

and sensitivity of human glioma cells to ganciclovir. Hum Gene Ther 6:1507-13

186